Biological Evolution: An Introduction 0521012058, 9780521012058

Biological evolution, the theory of natural selection and of common descent, is a triumph both of human reasoning and sc

523 140 8MB

English Pages 278 [281] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Biological Evolution: An Introduction
 0521012058, 9780521012058

Table of contents :
Copyright
Contents
Preface
Acknowledgements
1 Biological Evolution: The Beginnings of the Story
The Development of Evolution as a Science
The Years before Publication of Origin of Species
So, What Is Evolution?
Change and Species Formation
Natural History and Classification
Exploring the Development and Progress of Life on Earth
The Galapagos Islands and Darwin’s Finches: A Case Study
The Finches
Classification and the Galapagos Finches
Darwin’s Finches and the Origin of Species
The Galapagos Islands and Natural Selection
2 Reviewing the Evidence for Evolution
Homology and Comparative Anatomy
Embryology
Vestigial Organs
The Fossil Record
Fossils and Phylogeny
Biogeography
Observational and Experimental Evidence
3 Genetic Variation within Populations
Inheritance and Variation
Early Ideas Regarding the Continuity of Life
Biological Inheritance and the Work of Gregor Mendel
Mapping the Genome
Origins and Maintenance of Variation
Mutation
What Sorts of Genes Are Needed by Living Things?
Genotypic and Phenotypic Variation
Genes in Populations
Variation within Populations
Variation between Populations
Population Genetics
4 Natural Selection and Adaptive Change
Natural and Artificial Selection
Selection in Populations
Polymorphism
Heterozygote Advantage
Directional Selection and Local Adaptation
Sexual Selection
Genetic Drift and the Adaptive Landscape
The Unit of Selection
5 Evolution and Development
Evolutionary Developmental Biology (Evo-Devo)
The Epigenetic Landscape
Homeosis
Hox Genes
The Body Axes and Segmentation
The Dorsoventral Axis
Functional Analogy
The History of Hox Genes
The Divergence of Body Plans
Homeotic Genes and Control of Development in Higher Plants
Evolutionary Developmental Repatterning
6 The Origins of Biodiversity
Species Concepts
Isolating Mechanisms
Speciation
Speciation through Polyploidy
Parapatric Distribution, Speciation and Hybrid Zones
Sympatric Speciation
The Explosive Speciation of Cichlids
7 Taxonomy and the Diversity of Life
Linnaeus and Classification
Lamarck and the Scala Naturae
Classification and Evolution
Chasing Ancestors
Developing a Modern, Biological Classification
An Objective Classification?
Phenetics
Cladistics
Molecular Taxonomy
Nomenclature
Classification and Big Data
8 The History and Origins of Life on Earth
What Is Life: Characteristics of Living Things
Origins of Life
The First Organisms
Origins of the Eukaryotes and the Evolution of Sex
Multicellularity and the Higher Taxa
The Evolution of Animals
The Evolution of Plants
Movement onto Land
9 Molecules and Evolution
The Early Earth
Replication and the RNA World
Gene Trees
DNA and RNA Phylogenies
Rates of Molecular Evolution
Molecular Clocks
Phylogenomics and Transposable Elements
Lateral Gene Transfer
Genomics and ‘Big Science’
10 Human Evolution
Looking at Mammals
Becoming Human
Palaeobiology and the Human Lineage
Modern Humans
Evidence from the Human Genome
Human Success
Human Cultural Evolution
Are We Still Evolving?
11 Trends and Patterns in Evolution
Rates of Evolution
Measuring Rates of Evolutionary Change
Extinction and Patterns of Mass Extinction
Heterochrony and Life History Strategies
Are Trends in Evolution Progressive?
Biological Evolution As Science
12 Questions, Debate and Controversy
Questions in Evolutionary Biology
Life’s Continuing Existence
Evolution and Religion
From So Simple a Beginning
References
Recommended Reading
List of Figure Credits
List of Chapter Reviewers
Index

Citation preview

Biological Evolution An Introduction Biological evolution, the theory of natural selection and of common descent, is a triumph both of human reasoning and scientific undertaking. The biological discipline of evolution contains both a chronicle of human endeavour and the story of life on Earth. This book is concerned with living forms and how they developed from ‘simple and unpromising beginnings’. It considers evolution as both process and product. The author, an experienced teacher and educator, employs a historical narrative, used to convey the idea of ‘change with modification’ and to emphasise the relevance of evolution to contemporary bioscience. Biological evolution has now become part of the scientific orthodoxy, and this accessible text will assist undergraduate students in the biological sciences within any ongoing debate. Mike Cassidy is a Teaching Fellow in the School of Education at Durham University, UK. He has taught in schools, colleges and universities, and has co-authored advanced level Biology textbooks. He has worked extensively with the Royal Society of Biology and is a Fellow both of that society and the Linnaean Society.

Biological Evolution An Introduction MIKE CASSIDY Durham University

University Printing House, Cambridge CB2 8BS, United Kingdom One Liberty Plaza, 20th Floor, New York, NY 10006, USA 477 Williamstown Road, Port Melbourne, VIC 3207, Australia 314–321, 3rd Floor, Plot 3, Splendor Forum, Jasola District Centre, New Delhi – 110025, India 79 Anson Road, #06–04/06, Singapore 079906 Cambridge University Press is part of the University of Cambridge. It furthers the University’s mission by disseminating knowledge in the pursuit of education, learning, and research at the highest international levels of excellence. www.cambridge.org Information on this title: www.cambridge.org/9780521812689 DOI: 10.1017/9781139016018 © Mike Cassidy 2021 This publication is in copyright. Subject to statutory exception and to the provisions of relevant collective licensing agreements, no reproduction of any part may take place without the written permission of Cambridge University Press. First published 2021 Printed in the United Kingdom by TJ Books Ltd, Padstow Cornwall A catalogue record for this publication is available from the British Library. ISBN 978-0-521-81268-9 Hardback ISBN 978-0-521-01205-8 Paperback Cambridge University Press has no responsibility for the persistence or accuracy of URLs for external or third-party internet websites referred to in this publication and does not guarantee that any content on such websites is, or will remain, accurate or appropriate. The author and publisher have acknowledged the sources of copyright material where possible and are grateful for the permissions granted. While every effort has been made, it has not always been possible to identify the sources of all the material used, or to trace all copyright holders. We would appreciate any omissions being brought to our attention.

Contents

Preface Acknowledgements

page ix xi

1

Biological Evolution: The Beginnings of the Story The Development of Evolution as a Science The Years before Publication of Origin of Species So, What Is Evolution? Change and Species Formation Natural History and Classification Exploring the Development and Progress of Life on Earth The Galapagos Islands and Darwin’s Finches: A Case Study The Finches Classification and the Galapagos Finches Darwin’s Finches and the Origin of Species The Galapagos Islands and Natural Selection

1 1 3 4 6 6 8 10 13 15 20 22

2

Reviewing the Evidence for Evolution Homology and Comparative Anatomy Embryology Vestigial Organs The Fossil Record Fossils and Phylogeny Biogeography Observational and Experimental Evidence

25 25 29 32 35 38 41 43

3

Genetic Variation within Populations Inheritance and Variation Early Ideas Regarding the Continuity of Life Biological Inheritance and the Work of Gregor Mendel Mapping the Genome Origins and Maintenance of Variation Mutation What Sorts of Genes Are Needed by Living Things? Genotypic and Phenotypic Variation

45 45 50 50 53 56 57 61 63

vi

Contents

Genes in Populations Variation within Populations Variation between Populations Population Genetics

65 66 70 72

4

Natural Selection and Adaptive Change Natural and Artificial Selection Selection in Populations Polymorphism Heterozygote Advantage Directional Selection and Local Adaptation Sexual Selection Genetic Drift and the Adaptive Landscape The Unit of Selection

74 76 80 81 85 86 88 91 91

5

Evolution and Development Evolutionary Developmental Biology (Evo-Devo) The Epigenetic Landscape Homeosis Hox Genes The Body Axes and Segmentation The Dorsoventral Axis Functional Analogy The History of Hox Genes The Divergence of Body Plans Homeotic Genes and Control of Development in Higher Plants Evolutionary Developmental Repatterning

94 94 96 97 98 100 102 104 105 106 108 111

6

The Origins of Biodiversity Species Concepts Isolating Mechanisms Speciation Speciation through Polyploidy Parapatric Distribution, Speciation and Hybrid Zones Sympatric Speciation The Explosive Speciation of Cichlids

112 113 116 117 119 120 125 126

7

Taxonomy and the Diversity of Life Linnaeus and Classification Lamarck and the Scala Naturae Classification and Evolution Chasing Ancestors Developing a Modern, Biological Classification An Objective Classification?

130 131 133 135 136 138 139

Contents

vii

Phenetics Cladistics Molecular Taxonomy Nomenclature Classification and Big Data

139 142 147 149 152

8

The History and Origins of Life on Earth What Is Life: Characteristics of Living Things Origins of Life The First Organisms Origins of the Eukaryotes and the Evolution of Sex Multicellularity and the Higher Taxa The Evolution of Animals The Evolution of Plants Movement onto Land

153 154 156 160 161 164 168 177 179

9

Molecules and Evolution The Early Earth Replication and the RNA World Gene Trees DNA and RNA Phylogenies Rates of Molecular Evolution Molecular Clocks Phylogenomics and Transposable Elements Lateral Gene Transfer Genomics and ‘Big Science’

182 182 185 186 190 191 192 193 195 196

10

Human Evolution Looking at Mammals Becoming Human Palaeobiology and the Human Lineage Modern Humans Evidence from the Human Genome Human Success Human Cultural Evolution Are We Still Evolving?

198 199 202 206 211 212 215 219 221

11

Trends and Patterns in Evolution Rates of Evolution Measuring Rates of Evolutionary Change Extinction and Patterns of Mass Extinction Heterochrony and Life History Strategies Are Trends in Evolution Progressive? Biological Evolution As Science

223 224 226 230 234 237 238

viii

Contents

12

Questions, Debate and Controversy Questions in Evolutionary Biology Life’s Continuing Existence Evolution and Religion From So Simple a Beginning

241 241 246 247 250

References Recommended Reading List of Figure Credits List of Chapter Reviewers Index

251 255 262 264 265

Colour plates can be found between pages 148 and 149.

Preface

A textbook is more than a simple source account or provider of information. We live in an information age where factual description and scientific explanation are readily available on-line. And so, the textbook (particularly an introductory text such as this) should also convey ideas, stories, context and controversies as well as fulfilling its primary teaching role. Inevitably, in discussing biological evolution, there will be overlap between academic disciplines (genomics, molecular biology, history of science, palaeontology, anthropology and zoology) but the outcome remains the same – evolution provides both a profound and true account of life on Earth. Evolution is an accepted fact supported by overwhelming evidence. Rules of scientific evidence apply here in the same way as they do in molecular, physical or chemical studies; and we can note that recent research into genomics and molecular phylogenetics is continuing to yield new insight into biological evolution. History though, if anything, teaches us not to be complacent; scientific principles can be re-examined, repurposed and redefined. It is my hope that this text will inspire the reader to explore further the intricacies of biological evolution and ultimately to understand the origins of ourselves and the world around us. We learn effectively through stories. And the biological discipline of Evolution contains both a chronicle of human endeavour and the story of life on Earth. This book is concerned with living forms and how they developed from ‘simple and unpromising beginnings’. It considers evolution as both process and product. An historical narrative is employed; used to convey how the idea of ‘change with modification’ developed and what evolution now means to contemporary bioscience. The topic of Evolution is taught in schools, colleges and universities (it has also been included in the UK National Curriculum for Primary schools) and its central role in the study of the Life Sciences is now well understood. Evolution of course provides a unifying theme, a scaffold on which to place our developing understanding of past and present biota. In an age where students discuss the evolution of the cosmos or the evolution of the mobile phone it was thought appropriate here to use the more correct epithet, Biological Evolution as its title. Biological evolution, the theory of natural selection and of common descent, is a triumph both of human reasoning and scientific endeavour. And although, for most of us, the story begins with Charles Darwin and Alfred Russel Wallace in the mid-

x

Preface

nineteenth century, the idea of biological change over time was not new. Through primitive animism and the later philosophies of the ancient world, the history of evolutionary thought takes in several millennia and several different world views. The Age of Enlightenment, including the Scientific Revolution of the seventeenth and eighteenth centuries, also had a profound influence on evolutionary thought. Scientific advance may seem to appear rapidly in human history, but the reality is the ‘fine tuning’ of ideas and refinement of major concepts takes many generations. The same is true for biological evolution. The basic ideas of selection and modification were laid down by Darwin and Wallace, but over the subsequent 150 years or so new insight into both macro- and microevolutionary change has become evident. Not least of which are the syntheses of Evolution and Heredity together with Evolution and Development. The new sciences of genomics and bioinformatics are providing even further detail concerning the mechanism of change while advances in palaeontology, embryology, biogeography and geology yield yet more insight. Through context (how and when ideas were first formed) and through argument and debate the text will both encourage exploration and provide an explanation for evolution. This book is intended as an introduction to the subject of biological evolution for the undergraduate student of biology (along with students of anthropology, psychology, genetics and allied professions). Its text is intended to be both comprehensive and detailed where necessary, but it is hoped that the narrative style and historical context will also appeal to anyone with an eye for a good story. The book is structured in such a way as to introduce the main ideas initially and then explore details of mechanism (‘how evolution occurs’) and product (‘what has evolution produced’). As far as possible, detailed mathematical accounts and complex chemistry have been omitted. For convenience, important technical terms are written in bold type while in-text citation and references have been kept to a minimum with both a ‘References’ and a ‘Recommended Reading’ section at the end of the book. The book is arranged into 12 chapters. Early sections deal with a historical account of the major evolutionary figures and the evidence put forward to support their theories. The middle chapters look in detail at microevolutionary processes, while a ‘macro’ approach, the history, origins and progression of life on Earth, follow on. The final chapters on trends, debate and controversies explore recent advances in evolutionary science along with the cultural impact of biological evolution in the nineteenth, twentieth and twenty-first centuries. An analysis of this kind will inevitably explore the bigger issues of science and religion, communicating science and the misuse of scientific theory with evolution as its central theme. Biological evolution has now become part of the scientific orthodoxy, but it is not, of course, without its detractors. It is hoped that this text will assist students within the on-going debate.

Acknowledgements

The book owes much to past students, friends, family and colleagues, in particular initial conversations with Dr Alec Panchen. My own thinking on the ideas and concepts of evolution has also developed profoundly through tutoring the next generation of biologists and biology teachers both at Warwick and Durham Universities. I acknowledge, with gratitude, the help of staff at Cambridge University Press. Other colleagues have also assisted in reviewing chapters and allowing permission to use figures; their names (and my thanks) are found at the end of the book. The completion of this book has been influenced of course by my family, for without them it would not have been possible. My wife and sons have been steadfast in their support; to Jacquelyn, Oscar and Athol for their assistance, perseverance and insight. There are so many discerning books and inspiring communicators of evolutionary biology – I wouldn’t know where to begin thanking them. But I do know that the scientific establishment is so much the better for their presence, and I hope that this text will assist the next generation of life scientists in promoting their own endeavours.

1

Biological Evolution The Beginnings of the Story

This book is about scientific ideas and the evidence needed to exemplify and support the theory of evolution. It explores current biological diversity and asks the question how all the various life forms on our planet came about. Why do we have so many different species, and what processes cause biological change over geological time?

The Development of Evolution as a Science An evolutionary narrative is often thought to begin with Charles Darwin, but historically evolutionary ideas have been with us for at least two millennia. Classical Greek philosophers such as Theophrastus (371 BC–287 BC) and Aristotle (384 BC–322 BC) were keen naturalists providing some of the first direct observations and empirical accounts of the natural world. And just as Theophrastus was studying plants (he was the first to systematically group plants) in lagoons nearby, Aristotle was contemplating the essential differences between plants and animals. Aristotle was interested in boundaries between species; not that he was presupposing speciation – for Aristotle believed in the ‘Ladder of Life’, a fixity of animal forms, moving from worms and simple creatures, through stages, to fish etcetera with Man at the top superseded only by the Gods. But Aristotle was prescient in that he saw nature as ‘changeable’ (in the manner rivers change the landscape over time) and ‘graded’ (as animals vary both from one another and from other animals), but species he believed were immutable and unchanging. Regarding the origins of life, he disagreed with both Empedocles (490 BC–430 BC) who had earlier suggested that life arose through chance assemblages in some early primordial soup and Anaximander (610 BC–546 BC) who speculated that all life arose in water. Charles Darwin himself thought Aristotle to be a proto evolutionist (not surprising as he was an acute observer of nature and keen to remove mysticism from the debate). But he was mistaken on this count due to an error made by a local town clerk who had mistranslated Aristotle’s ‘physics’. Aristotle was not supporting any species change but rebutting the argument put forward earlier by Empedocles. Darwin was not a classicist! Later, as the classical texts of the Middle Ages gave way to the European Renaissance (fourteenth to seventeenth centuries) and then to the ‘Age of Enlightenment’ (eighteenth century), a profound shift in thinking was taking place. Encouraged by voyages of discovery around the world, wealthy individuals began to collect attractive

2

Biological Evolution: The Beginnings of the Story

and interesting specimens and display these within their ‘cabinets of curiosities’. Notably, the collection of Sir Hans Sloane became the basis of the collections now contained within the British Museum. Similarly, improvements in the technology of observation (telescopes and microscopes) together with developments in mass communication printing provided a further impetus for human intellectual voyages of discovery and with it the popularisation of science. The process of collecting, cataloguing and displaying specimens eventually developed into a much more systematic endeavour. Collections of minerals and biological specimens were described and organised to uncover underlying organising principles. Explanations were also sought for the observations now being made. In truth, a scientific revolution was taking place where myth was to be replaced by theory, conjecture with evidence and simple curiosity with systematic investigation. Francis Bacon’s empirical approach led ultimately to ‘new ways of knowing’. Classical thinkers of the Middle Ages had been overtaken by what is referred to as the natural philosophers of the seventeenth and eighteenth centuries. Natural philosophy was a thoughtful and systematic study of the natural world. Subsequently the ‘scientist’ (a new term coined by Thomas Whewell in the mid-nineteenth century) would be associated with a practice involving hypothesis formation and rigorous testing of ideas. Charles Darwin (1809–1882) of course was an inspired scientist. Darwin’s observations on biological complexity were systematic and his explanation of how this complexity arose took the form of a carefully reasoned argument. He used evidence to support his claims; evidence that could be checked and replicated by the wider scientific community. The earlier world views of Newton, Leibniz and Hobbes provided a rigid, almost clockwork view of the world, whereas in the midnineteenth century a more historical thinking prevailed. Examples of this new mind set include political thinkers such as Marx and Hegel who employed a dynamic and historical view of world events. Their thinking relates to a view of the world changing not the fixed view of their predecessors. Darwin’s half-cousin Francis Galton (1822–1911) had already explored increases in human population and its potential consequences while his grandfather, Erasmus Darwin (1731–1802), a prominent poet and biologist, alluded to a process of evolution and biological change in two of his long poems. Contemporary with Charles Darwin, nineteenth century geologists such as Charles Lyell (1797–1875) and Adam Sedgewick (1785–1873) emphasised that the planet too was not a fixed entity but had undergone profound change ‘throughout the long expanse of history’. Limestone rock strata scattered throughout the British Isles demonstrated that these locations were once shallow seas with teeming marine life and not the Southern uplands and Yorkshire dales scenery that we see now. The seventeenth-century image of an unmoving, static world was slowly being replaced by a more dynamic perspective. In the early nineteenth century, following the French revolution, there was a break with the more ‘classical’ approach. And proponents such as Lamarck and Saint-Hilaire challenged the (by now becoming outdated notion) of the ‘fixity of species’. This mind set affected Charles Darwin in his attempts to understand biological complexity. In 1859 Darwin published his On the Origin of

The Years before Publication of Origin of Species

3

Species by Means of Natural Selection together with his own description of biological change – ‘descent with modification’. Darwin also included a means by which these events could be explained, ‘natural selection’. Charles Darwin was both a product of and contributor to this new way of thinking (or paradigm shift as Thomas Kuhn [1996] later called it). As ‘natural philosophy’ gave way to ‘natural science’, a more rigorous, experimental approach or scientific method began to define scientific endeavour. Individuals such as Francis Bacon, 1561–1626 (philosopher, parliamentarian and scientist), Michael Faraday, 1791–1867 (the most eminent experimental chemist of his day) and William Whewell, 1794–1866 (President of the Geological Society) exemplified this approach. Whewell was a source of inspiration for Charles Darwin. Later that century biological science (the term ‘biology’ was coined in 1800 in an obscure German footnote) developed concepts such as the cell theory, principles of homeostatic control and impressive advances in animal and plant physiology through rigorous observation and experimentation. Biological evolution was slightly different, however. It did not at that time employ experimentation, but rather a systematic collection of evidence to answer questions together with an acutely reasoned argument. Following its synthesis with twentieth-century genetics, biological evolution rapidly became the cornerstone of biology; as Theodosius Dobzhansky famously says in his 1973 essay, ‘Nothing in biology makes sense except in the light of evolution’. The history of evolution as an idea has had a long gestation, at times controversial, continuing in the twentieth century with development of evolutionary genetics. Genomics, a subject that did not exist before the twenty-first century, heralds a new chapter in our understanding.

The Years before Publication of Origin of Species The year 1830, like many of those in the previous four decades, had been a turbulent one in French history. There had been revolution in Paris and the King was forced to abdicate. So when a friend called on the German poet Johann Wolfgang von Goethe in Weimar, he was prepared to agree that a great explosion had taken place in European affairs. But he was flabbergasted to discover that Goethe was referring not to French political upsets but to an acrimonious debate between two of the most noted comparative anatomists of the day, Georges Cuvier and Étienne Geoffroy Saint-Hilaire. For Goethe too was a considerable anatomist and appreciated the significance of the event. The debate between the two former friends and current colleagues was not about evolution. The question, debated before a noisy audience in the premises of the Académie de Sciences in Paris, was about the correct way to interpret anatomical resemblances between different species of animals. To Cuvier, identity of structure meant identity of function; an animal, any animal, remained alive because it functioned like a well-coordinated machine. Every characteristic, internal and external, was created to serve its current way of life – no further explanation was required.

4

Biological Evolution: The Beginnings of the Story

Geoffroy Saint-Hilaire agreed that functional anatomy was a worthy study, that anatomical features subserved a vital function. But to him functional anatomy was not a complete explanation. Quite apart from their function, the anatomical features suggested variation on an underlying plan. The proper task of ‘philosophical anatomy’ was to elucidate that plan – what, apart from their various ways of life, did all vertebrate animals have in common: could one reconstruct a basic vertebrate animal? Over the years, Geoffroy Saint-Hilaire had attempted to implement this programme to the increasing irritation of Cuvier, but when Geoffroy Saint-Hilaire suggested that invertebrate animals, such as insects, lobsters and molluscs, also shared the same plans as vertebrates, open disagreement broke out. Cuvier was a student of adaptation (that is the machine-like coordination of animal parts and the ‘fit’ of the whole animal to its environment), while Geoffroy was a student of homology (resemblances between species reflecting a common plan). Homology does not necessarily imply common ancestry, but it was due to the genius of Charles Darwin, through his Origin of Species, published in 1859, that both aspects of comparative biology were combined into a successful theory of evolution (Darwin, 1859).

So, What Is Evolution? What do we mean by the term ‘evolution’? There are several different interpretations. Originally evolution implied some sort of unfolding, like the opening of a flower (Latin = evolutio: an unrolling), but latterly it has acquired a wider meaning, implying a general process of change. Darwin’s phrase ‘descent with modification’ accurately describes the process of biological change. This book is about biological (or organic) evolution – a system of theories put forward to explain both diversity and the relationships between different types of living thing. If we wish to understand the theory of evolution, we need to consider the answer through a series of subordinate questions. A theory is an established idea or organising principle used to explain a body of information. It covers a wide range of facts and forms and is said to possess both explanatory and predictive power. A theory is more than just mere speculation; a theory is a precise conceptual framework that supports the data. The theory of evolution by natural selection is a powerful explanatory tool. It makes predictions such as the existence of genetic variation (otherwise evolution could not happen) and patterns of speciation found in fossils (as seen in rock strata). It is supported by evidence from a range of sources, palaeontological, genetic, anatomical, behavioural and biogeographical; it even supports what Coyne (2009, in his book Why Evolution Is True) refers to as retrodictions, facts and data that ‘make sense only in the light of the theory of evolution’. In the construction of any theory there are two component parts: 1. the data to be explained (in philosophical terms we call this the explanandum) and 2. the theory or the explanation itself (the explanans).

So, What Is Evolution?

5

So, what does evolution attempt to explain, what is its explanandum? Several answers have been offered to this question but there is a difference of emphasis among experts. Here are some possibilities. The explanandum – evolution attempts to explain: 1. Why there are a staggering number of different types of living things alive on Earth today (some 30 million possible species)? 2. How it is possible to classify organisms in a hierarchical grouping, in Darwin’s phrase ‘in groups within groups’. Is there something real about biological classification? Does it suggest genuine relationships? 3. How the fossil record chronicles the biota – a sum of all life forms over time. 4. Why organisms appear to be particularly well adapted to their environment. From these four questions above stem different schools of evolutionary research. And in order to answer the four questions above we can suggest, The explanans (In the same order as the questions were posed these are): 1. Those wishing to explain biodiversity and the ‘staggering number of different types of living things’ are likely to be interested in speciation; the division, in time, of one species into two or more and the mechanisms by which this occurs. 2. Taxonomists, interested in the classification and the hierarchical grouping of organisms, are concerned not only with constructing classifications but also with reconstructing the history of life (to which others including palaeontologists and molecular biologists also contribute). 3. Palaeontologists study fossils and explore life forms in different geological periods and can comment upon rates of evolution. 4. It is probable that most evolutionary biologists are preoccupied with the origin of adaptations – the reasons why adaptation is adequate rather than perfect and whether all the characteristics of organisms should be explained by natural selection. To answer our question therefore (so, what is evolution?) we might say that evolution is a process of biological change – a theory that attempts to explain biodiversity together with an explanation in terms of differential reproductive success. In addition to these lines of research there is a newly important branch of evolutionary theory, that of the evolution of development (or ‘Evo-Devo’ as it is known to its practitioners). For many reasons current evolutionary ideas do not fully explain how the development of individual organisms evolved. But in recent years there has been an explosion of knowledge in the role of the genome in animal development and the application of this knowledge to evolutionary problems. It should be clear from what has been said so far that not only are there several sets of data that can be explained by evolutionary theory, but there are also several types of explanation. Together these represent the multifaceted discipline of evolutionary biology.

6

Biological Evolution: The Beginnings of the Story

Change and Species Formation In studying evolution, one is inevitably exploring biological change, the formation of new species together with the extinguishing of others. But change and dynamism appear to be features of the world in which we live. Over its four billion or so year history the Earth has undergone profound change in terms of its geology, its atmosphere, the landscape, the climate and its constituent biota. Indeed, change in the abiotic (non-living) world often precedes or even dictates change in the biotic. Further proof, if needed, that all aspects of the natural world are interwoven Perhaps a more cogent argument arises when scientists look beyond our own planet for signs of life. This new science of exobiology (also referred to as Astrobiology) needs to consider how extraterrestrial life might present itself. It presumably will need to secure an energy source and it will need to carry out various processes including coordinated activity and reproduction, but importantly (for the argument presented here) life will be seen to evolve. Evolution, or hereditable biological change over time, is now generally seen as one of the handful or so major characteristics of living things. Professor Gerald Joyce at the Salk Institute in the United States is an astrobiologist and an expert in the field of in vitro evolution (recreating the biomolecules of early life). Perhaps he has provided us with the best definition of life: A self-sustaining chemical system capable of Darwinian evolution

One of Charles Darwin’s greatest achievements is to suggest a mechanism for the observed biological change over time – and that is natural selection. His theory of natural selection is both simple and elegant. Yet it is not reducible to the conventional rules of physics and chemistry. In this respect the biological sciences may be considered as inhabiting two epistemological ‘spaces’; on the one hand, the sciences of genetics, physiology, medicine and neuroscience (disciplines that are reducible to physical laws) and on the other, behaviour, community ecology and evolution which are not. Evolution it is argued belongs to this latter branch of whole organism biology where possible emergent properties arise and different research paradigms are needed.

Natural History and Classification Organising our knowledge of the natural world and naming objects is a characteristic of human societies. Allied to this peculiarly human activity is the search for order and a desire to explain the world as it appears to us. The biological discipline dealing with the classification or grouping of organisms is known as taxonomy; this forms part of a more general speciality known as systematics (a study of the types and diversity of organisms). Confusingly, some biologists – mostly botanists – refer to a classification as a ‘taxonomy’. Nomenclature (the naming of organisms) is a highly prescribed business. Before organisms can be classified, it is essential to have an agreed naming system.

Natural History and Classification

7

This applies not only to the naming of species but because classification of organisms is always expressed as a hierarchical structure (‘groups within groups’), there must also be rules about the naming of higher ranks. The whole system is policed by various International Commissions, most notably one for Zoology and one for Botany. Until the early 1960s methods for classifying organisms were ill-defined despite the fact that systematists claimed they were producing ‘evolutionary classifications’. Methods were largely intuitive. But then there arose not one, but two methods of classification, both claiming to be uniquely objective. They are known as phenetics (originally called ‘numerical taxonomy‘) and cladistics. Their practitioners often became bitter rivals, while both poured scorn on the easy-going and intuitive evolutionary taxonomists. The dust has now settled, and methods related to both phenetics and cladistics are in use for different taxonomic purposes. Natural history as an academic enterprise has a long and distinguished history in the United Kingdom. The oldest biological society in the world, The Linnaean Society of London, was founded in 1788 to honour the botanist (Carl Linnaeus), his works and his legacy – his efforts in systematising the living world. Elsewhere in Britain natural history became more organised with the standard works on identification produced. These included John Ray’s Catalogus Plantarum Angliae and Martin Lister’s Historiae Animalium Angliae, both published in 1678. It was in Plant Science or Botany that the discipline of natural history was first formalised. This is not surprising given the relevance of plants and plant products to the early study of medicine. The Society of Apothecaries based in London not only initiated the famous Physic garden at Chelsea but also promoted field trips into the local countryside. The earliest of these excursions was in May 1620 (the date of the voyage of the Mayflower to the New World). The Aurelians, as the lepidopterists (butterfly hunters) of the day like to call themselves, were another early specialist society. In the mid-eighteenth century, natural history was more of a fashionable subject than a scientific one. It was perhaps the Victorians in the nineteenth century who forged natural philosophy to become the precursor of the more academic disciplines of Biology and Geology. Charles Darwin’s seminal work (Origin of Species) in 1859 interestingly provided a unifying theory for both the plant and animal sciences. In 1866 a Chair in Zoology and Comparative Anatomy was created at Cambridge University, and the Education Act of 1870 brought a breakthrough in the teaching of Elementary Science. Indeed, there was such a shortage of teachers that the eminent zoologist Thomas Henry Huxley was asked by the government to set about providing a ‘crash course’ for teachers in botany and zoology. There are many clubs, associations and learned societies that have contributed to our knowledge of the natural world. Both amateur and professional biologists are employed in the study of flora and fauna, local and national. It is upon this knowledge base, prepared by the natural historian, that the modern disciplines of taxonomy, ecology, ethology and (ultimately) evolutionary biology are founded. An early example of a natural historian exploring evolutionary theory is that provided by Canon Henry Baker Tristram, born in 1822. ‘The great Gun of Durham’,

8

Biological Evolution: The Beginnings of the Story

as he was known, was an authority on birds in Durham, Northumberland and Palestine. As president of the British Association and Canon of Durham University’s College, Tristram (described as ‘a close observer and diligent collector’) was one of the first people to accept, in print, Darwin’s theory of evolution. This he did in an article in 1859 (less than one year after the publication of Origin of Species) in the ‘Ornithology of North Africa’: Writing with a series of about 100 Larks of various species from the Sahara before me, I cannot help feeling convinced of the truth of the views set forth by Messrs. Darwin and Wallace in their communication to the Linnaean Society . . . it is hardly possible I should think to illustrate this theory better than by the Larks and Chats of North Africa. (The Ibis, Volume 1, 1859)

Tristram then proceeds to discuss ‘gradual modifications of colouration and anatomical structure’ where ‘in the struggle for life . . . a very slight change for the better . . . would give the variety that possessed it a decided advantage over the typical or other forms of the species’ (Tristram, 1859: pp. 429–430). These views were also expressed in his Presidential address to the Tyneside Naturalists Field Club. This was a brave act coming from an Anglican churchman, but indicative of the growing acceptance of evolutionary theory.

Exploring the Development and Progress of Life on Earth Reconstructing the history of life is usually regarded as the task of evolutionary biologists in general and palaeontologists, whose discipline takes in aspects of both biology and geology. Essentially, palaeontologists collect and prepare (that is clean up) fossils and then try to make valid statements about the anatomy, ecology and even behaviour of the organisms their specimens represent. Most palaeontologists are taxonomists and attempt to say something about the historical significance of their fossils by including them in a classification that also embraces living species. A further category of evolutionary biology is that of the ‘adaptationists’ (there does not seem to be a suitable collective noun). Many are particularly interested in the evolution of behaviour (including human behaviour) and term themselves ‘behavioural ecologists’ or ‘sociobiologists’. Their principal preoccupation is with testing or applying Darwin’s theory of natural selection to the anatomy, behaviour and ecology of animals. One thing Darwin could not do was provide a valid account of heredity – the mechanisms by which the characteristics of one generation are passed on to succeeding generations. No one could blame him for that as the work of Gregor Mendel (and hence the beginning of modern genetics) was only ‘rediscovered’ in the year 1900. At first a number of scientists believed that Mendel’s conclusions refuted Darwin’s theory of natural selection. The two theories were happily reconciled in the late 1930s and early 1940s in the so-called ‘Synthetic Theory’ of evolution. This new synthesis (the Synthetic Theory or Modern Synthesis) proposed that variation was brought about by random events and that populations evolve by means of changes in gene

Exploring the Development and Progress of Life on Earth

9

frequency (e.g. those brought about by natural selection). The Synthetic Theory is sometimes called ‘Neo-Darwinism’, the revival of an older term with a somewhat different meaning. Evolutionary change can occur both above and below the level of species. Genetic change within a population, or below the species level is referred to as microevolution (that is changes in gene frequencies, mutation etc.). It is possible to demonstrate microevolution. Macroevolution, on the other hand, is evolution above the species level, including speciation. Its phenotypic changes affect the lineage of organisms and the ultimate appearance of higher groups (for example, the evolution of insects and the appearance of land plants). Macroevolution takes place over a much larger time scale and its progress is inferred using various lines of evidence, fossil appearance, radiometric dating, chemical analysis and degrees of relatedness. By the mid-1960s it became possible to study evolution at the molecular level. In studying proteins, it became apparent that there was a greater diversity of molecular form within populations than previously imagined. Techniques such as gel electrophoresis confirmed the amino acid sequences of these molecules, while rates of change led to the suggestion of the possibility of ‘molecular clocks’. Motoo Kimura (1924–1994), a Japanese population biologist, hoped to combine the discipline of population genetics with the newly emerging molecular data. What emerged was a realisation that the observed variation within groups was too large to be explained simply by natural selection. He therefore proposed an alternative hypothesis, that of the Neutral Theory of Evolution. In this he postulated that molecular evolution was driven not necessarily by Darwinian natural selection but by random, non-adaptive changes within the genome. Results of molecular studies have proved to be increasingly important in understanding the evolution of life on Earth, while the neutralist–selectionist debate has proved to be a useful focus for studies of molecular evolution. To summarise, therefore, the Earth is a rationally ordered physical and biological system in which changes occur. In the mid-seventeenth century James Ussher, the archbishop of Armagh, stated that the Earth was created the night before Sunday 23 October in the year 4004 BC! He did this by carefully measuring biblical genealogies. By 1800, however, geologists had demonstrated that the Earth must be older (for instance by calculating the length of time it takes for an object with the mass of the Earth to cool down). And Darwin, like his mentor the geologist Charles Lyell, believed in the Principle of Uniformitarianism (an agreement that processes we see in the present day also occurred in pretty much the same way as they did in the past); both Darwin and Lyell believed in a continuous, gradual geological change. The continuity of geological events on Earth is mirrored by Darwin’s thoughts on organic evolution – a classic expression of this Principle of Continuity. Famously, in 1831 her Majesty’s ship ‘Beagle’ sailed from Devonport with the young naturalist Charles Darwin on board. And, as we now know, studies on the habits of the cuckoo, extinct quadrupeds, distribution of land shells and birds of the Galapagos Archipelago all contributed to his landmark text Origin of Species some 30 years later.

10

Biological Evolution: The Beginnings of the Story

The Galapagos Islands and Darwin’s Finches: A Case Study The Beagle’s orders were to survey and map the coastline of southern South America, then, following the Galapagos visit, to sail west via Tahiti, New Zealand and Australia, making astronomical and other observations. Darwin’s brief was, as guest naturalist, to study the geology and natural history. He landed home at Falmouth on 2 October 1836, nearly five years after the Beagle’s departure. Darwin recorded that ‘in July (1837) I opened my first notebook for facts in relation to the Origin of Species, about which I had long reflected, and never ceased working on for the next twenty years’. His great work on evolution, On the Origin of Species by Means of Natural Selection, or the Preservation of Favoured Races in the Struggle for Life, was not published until 1859. The Galapagos Islands, over the many years since Darwin’s visit, have acquired an almost mythical status in accounts of the development of his theory. Some popular myths have Darwin’s conversion to ‘transmutation’ (i.e. evolutionary change) occurring suddenly during his five-week stay on the Galapagos, but there is no evidence of this other than an ambiguous note written as he prepared a catalogue of his bird specimens from previous ornithological notes, nine months after leaving the Galapagos. He was referring to the mockingbirds (Mimus parvulus) collected from four of the islands: the specimens from Chatham and Albemarle he says appear to be the same, but the other two are different. On each island each kind is exclusively found; habits of all are indistinguishable. When I see these islands in sight of each other, and possessed of but a scanty stock of animals, tenanted by these birds, but slightly differing in structure and filling the same place in Nature, I must suspect that they are only varieties. If there is the slightest foundation for these remarks the zoology of Archipelagos – will be well worth examining, for such facts would undermine the stability of Species.

Darwin had also been told by the English vice governor of the Galapagos that the giant tortoises (Chelonoidis nigra) differed consistently from island to island but took little notice and did not collect museum specimens of the tortoises while there. The only tortoises collected by anyone (except for two babies kept as pets) were eaten by the Beagle crew and the skulls thrown overboard! Indeed, because of their size, hardiness and longevity, the tortoise population on the islands would be decimated by pirates and whalers who embarked onto the islands for shelter and provisions. It is reckoned that more than 100 000 of these lumbering reptiles (the megafauna of the Galapagos) were removed by seafarers. The Galapagos archipelago comprises 16 volcanic islands of differing ages with varying landscapes (Figure 1.1). The younger islands like Fernandina in the West are inhospitable with harsh, arid landscapes of volcanic ash and lava flows and little vegetation. The older islands like Santa Cruz to the East are clothed in vegetation and are the centre of the Galapagos’ famed biodiversity. The oldest islands like Espanola, around 4 million years old, are sinking into the ocean with erosion reducing the landscape to a flattened coastal remnant. The significance of this is that the diversity in

The Galapagos Islands and Darwin’s Finches: A Case Study

11

Figure 1.1 Sketch map of the Galapagos group of islands indicating the main sites referred to in

the text. Credit vasosh / iStock / Getty Images Plus.

animal and plant life is mirrored by the diversity in habitat. And since emerging from the Pacific Ocean ‘hot spot’ around four and a half million years ago, organisms have populated these islands (arriving from the South American coast) providing a unique insight into adaptive radiation and the evolutionary process. Radioactive dating has shown that the oldest islands are about 5 million years old, the youngest about half a million. The importance of the isolation of the Galapagos

12

Biological Evolution: The Beginnings of the Story

from the South American mainland to the study of Darwin’s finches is that it seems there was never any land connection between the two areas. If a common ancestral species of all Darwin’s finches arrived from South America, it must have done so by chance – on floating mats of vegetation (which still detach themselves from coastal Ecuador today), by being blown off course or by other infrequent means. Thus, the founding population of the ancestral Galapagos finches presumably consisted of a small number of individual birds, with no doubt many casualties en route. The environment of the Galapagos at first appears to the observer to be very inhospitable. Darwin’s first impression of ‘Chatham’ was not favourable: Nothing could be less inviting than the first appearance. A broken field of black Basaltic lava is everywhere covered by a stunted brushwood which shows little signs of life. The dry and parched surface, having been heated by the noonday Sun, gave the air a close and sultry feeling, like that from a stove: we fancied even the bushes smelt unpleasantly. (The Voyage of the Beagle, 1845, Chapter XVII, ‘Galapagos archipelago’)

Similarly, David Lack (1910–1973) speaks of ‘miles of dreary greyish brown thorn bush, in most parts dense, but sparser where there had been a more recent lava flow, and the ground still resembled a slag heap’. And yet the Galapagos have an enormously rich fauna, notably of birds, with many endemic species (unique to the island), and an equally rich marine life. A clue to the reason is the unique presence on the islands of a bird, whose group is more usually associated with Antarctica, the Galapagos penguin (Spheniscus mendiculus), the only equatorial penguin species. For islands on the equator the climate is often surprisingly cool and the sea conspicuously so. The explanation is the Humboldt Current which brings plankton-rich cold sea water up from the coast of Peru. The Galapagos climate is markedly seasonal. From roughly July to December, the Humboldt Current dominates, lowland air temperatures are cool and the rainfall slight. During this season, however, an inversion layer is created, and those islands with considerable highlands have those highland regions continuously wet, so they are covered in a rich green layer of plants. From January to June there is a warm, wet season in the lowlands with mostly clear skies but occasional heavy showers. Before moving on to talk in detail about those finches, there is one more important point to be made. The birds are not quite unique to the Galapagos Islands. About 300 miles (500 km) southwest of Costa Rica in Central America, and nearly twice that distance northeast of the Galapagos, there is a small (47 km2) island called Cocos (situated on the Cocos plate). Like the Galapagos, it is a volcanic island, but in other respects it is very different. There is no seasonality: the whole island experiences heavy rainfall throughout the year and consequently is covered by dense rain forest. There are only four resident land birds on Cocos Island. One of those is a ‘Darwin’s finch’! The vice governor’s comments about each island having its own type of tortoise extends also to other animals such as lava lizards, birds and various land snails. Each of these animals appears to have distinct forms unique to the different islands. Thus, one can postulate that the ancestral mainland ‘castaways’ that first arrived settled the

The Finches

13

various islands and therein adapted to their own unique environmental circumstances. So, tortoises on islands with lush vegetation developed high domed shells to move more easily (tracks made by tortoises in this dense vegetation are easily seen from the air), whereas those tortoises living where vegetation is poor acquired long necks and peaks at the front of their shells to allow them to reach up to the dominant plant (in this case the prickly pear cactus). In a similar vein, land snails in arid areas developed a small mouth (to avoid excessive water loss) and long, conical shells, while those snails in wet areas acquired wide mouths and globular shells. All of this points to a phenotypic plasticity and a rapid radiation of within-species forms adapted to its own environment and providing the potential for future speciation. But perhaps the most iconic example of island radiation and brisk speciation is the finches. Current reckoning is that there are 14 species of finches on the Galapagos. Darwin landed on four of the islands during his five-week stay (although he saw many more as the Beagle criss-crossed the archipelago). He collected specimens from the four islands but attached so little importance to inter-island variation that he mixed up the specimens from the first two islands (Chatham and Charles) that he visited. Darwin did not even recognise that all the species formed a closely related group until the Beagle specimens had been studied by the Zoological Society’s ornithologist John Gould. Gould explained to Darwin that all the Galapagos finches in the Beagle collection formed a group of species more closely related to each other than to any other birds. Darwin’s conversion to transmutation seems to have occurred soon after this explanatory meeting with Gould in mid-March 1837. Nevertheless, there is no account of the finches in the Origin of Species of 1859, although Darwin does say of them in the second edition of his account of the voyage (1845), ‘One might really fancy that from an original paucity of birds in this archipelago, one species has been taken and modified for different ends’. The Galapagos finches are small and mostly dull-coloured birds, little more than a sparrow size. The most striking feature of the whole group is the variation in the size and shape of the beak. The large ground finch Geospiza magnirostris has a massive parrot-like beak and a large heavily muscled head to support it. At the opposite extreme the little warbler finch, Certhidae olivacera, has a slender probing beak like that of a warbler! (Figure 1.2). Amazingly it took until the 1940s for general agreement that these beak shapes were adaptive, despite numerous early studies. This agreement resulted from the publication of an important book by David Lack, which popularised the phrase Darwin’s Finches as its title (Lack, 1947).

The Finches In the classification of organisms, the species is regarded as the fundamental taxonomic unit, and definitions of species (the ‘species concept’) have been a matter of vigorous debate since before Darwin’s time. Thus, any evidence bearing on the origin of a new species (or better still a whole series of new species) would be cogent evidence for evolution. The species concept is discussed in Chapter 6, but in general terms species are groups of freely interbreeding individuals separated from other

14

Biological Evolution: The Beginnings of the Story

1

2

4

3

1. Geospiza magnirostris. 3. Geospiza parvula.

2. Geospiza fortis. 4. Certhidea olivasea.

Figure 1.2 Woodcut of Darwin’s finches. As drawn by John Gould.

groups by failure to interbreed. In a classification species are gathered together into genera (singular genus), but genera are to some extent arbitrary and subject to personal taste; there is little discussion, if any, about ‘the genus concept’. It is thought that there are 14 species of Darwin’s finches on the Galapagos Islands, plus the one on Cocos Island. Not surprisingly large islands tend to have more resident species than smaller ones. Opinions vary as to the number of genera that should be used to group the finch species, but six genera are a generally acceptable number. They are as follows: 1. Geospiza: the ground finches o Large ground finch – Geospiza magnirostres o Medium ground finch – G. fortis o Small ground finch – G. fuliginosa o Sharp-beaked finch – G. difficilis o Large cactus finch – G. conirostris o Small cactus finch – G. scandeus There is strong evidence, and general agreement, that the ground finches form a natural group. All are related to one another more closely than to any other species of Darwin’s finch. Technically the genus Geospiza is a monophyletic group, that is, the group consists of ancestral species and all their descendants. 2. Camarhynchus: the tree finches o Large tree finch – Camarhynchus psittacula o Medium tree finch – Camarhynchus pauper o Small tree finch – Camarhynchus parvulus

Classification and the Galapagos Finches

15

The tree finches, as their name suggests, live much more in trees than do the ground finches. The tree finches are found more in the highlands and feed largely on insects. 3. Cactospiza: the woodpecker finches o Woodpecker finch – Cactospiza pallida o Mangrove finch – Cactospiza heliobates The woodpecker finch is famous for not only using tools, but also for modifying them to its purpose. The behaviour was discovered in 1919 and has been observed frequently since. The bird uses a cactus spine, or a twig, often broken off by the bird itself, as a probe held lengthwise in the beak to winkle out grubs, etc. from cracks in bark. It also climbs up and down vertically like a woodpecker. The mangrove finch has also been observed in tool use. The former is a mostly highland species, the latter is found only in the mangrove swamps of Isabela. 4. Platyspiza crassirostris: the vegetarian finch The only species in its genus; it lives in trees mostly in the highlands. Feeds on fruit, leaves and buds. 5. Certhidea: the warbler finches o The green warbler-finch – Certhidea olivacea o The grey warbler-finch – Certhidea fasca The warbler finches feed mostly on insects, even in flight. Certhidea olivacea is found in the central, high islands (Santa Cruz and Santiago). C. fusca is known to occur in four lower more peripheral islands. Darwin did not accept that the warbler finch was any close relation of the others until convinced by Gould. 6. Pinaroloxias inornata: the Cocos finch The single species from Cocos Island is small (13 g) with a slender, slightly curved beak like that of the warbler finch. Despite uniformity in appearance, a study has shown that individual birds are specialist feeders with a large variety of ways of life in the rain forest. Specialities include searching for insects in leaves and in branches, looking for crickets and grasshoppers among dead leaves and collecting nectar.

Classification and the Galapagos Finches Before any scientist can study objects, or phenomena, they must organise them some way. Ever since the days of the ancient Greeks, the most useful way of doing this has been to produce a hierarchical classification. Hierarchies come in two principal sorts. Both share the feature that they are defined by a series of ranks. The first is an exclusive hierarchy; an example here might be military rank. The second an inclusive hierarchy such as a taxonomic rank.

16

Biological Evolution: The Beginnings of the Story

The term ‘rank’ perhaps needs explanation; in the words of the evolutionary biologist Ernst Mayr (1904–2005): ‘Military ranks from private, corporal, sergeant, lieutenant, captain up to general are a typical example of exclusive hierarchy. A lower rank is not a subdivision of a higher rank; thus lieutenants are not a subdivision of captains’. In contrast though, a biological classification is an example of inclusive hierarchy. With Darwin’s finches, a single genus usually (but not always) includes several species. Then genera are grouped together in families and so on, at an ever more inclusive series of ranks (a set of nested groups). Biological classifications are usually also irregular – as an example, some genera have many species, others only one. By convention the hierarchy is also divergent; no species can belong to more than one genus. In his book Darwin’s Finches (1947), David Lack not only presents a written classification of the birds (with discussion of the priority of the whole grouping within higher ranks) but also draws up a diagram looking like a family tree, with individual species at the end of each branch. He describes this diagram as ‘an evolutionary tree’, thus drawing an important theoretical conclusion, which is by producing a (correct) classification one is producing not just a diagram of that classification but also of the pattern of descent: an inclusive irregular classification is best explained as a phylogeny. Lack drew on previous work in drawing up his diagram; his classification was principally based on not only comparison of the appearance of the birds, but also of their ecology and behaviour such as song. If, however, one could use some completely different method and different data to classify Darwin’s finches and the new method produced the same result, then one could feel that the classification was ‘correct’. It would also show that the classification was in some way real and not just a convenient grouping of data. A ‘real’ classification based on natural groupings (birds, fish, insects, etc.) has been referred to as a natural classification, whereas a classification of convenience (all the waterfowl, all the yellow flowers) has been called an artificial classification. Since Lack’s time, there have been numerous studies of Darwin’s finches, notably those of Robert Bowman from the 1960s to the 1980s including studies of beak function and song, which are outstanding not only as a series of works on the finches, but also as a thorough study of the ecology and evolution of animals. We can also note the research of husband and wife team Peter and Rosemary Grant (with a succession of colleagues, assistants and research students) from the 1970s to the present day. In recent years the finches have been reclassified using techniques derived from biochemistry and molecular biology. In the 1970s and 1980s attempts were made at reclassification using the electrophoresis of proteins. But this method in general was not able to distinguish between individual species in such a closely related group of birds. With the 1990s came the use of DNA sequencing. Deoxyribonucleic acid (DNA) is the genetic material in all animals, present in the nucleus of all cells as the famous double helix, but also present as a single strand in the numerous mitochondria, the tiny power-houses of the cell, scattered through the cell cytoplasm. The genetic code itself consists of four bases (A: adenine, C: cytosine, G: guanine,

Classification and the Galapagos Finches

17

T: thymine), an alphabet of four letters, anchored along the DNA molecule. But if two species of birds (or other organisms) diverge from one another over evolutionary time, point mutations can occur so that in one or both cases, one base is substituted for another at any site. The longer that two bird species have had a separate history, the more mutations are likely to occur, so that mutation number becomes a measure of elapsed time. Both Galapagos finches and Hawaiian honeycreepers have been used to study adaptive radiation. Honeycreepers (see Figure 1.3) such as Apane (Himatione sanguinea), L’iwi (Vestiaria coccinea), Amakihi (Hemignathus virens), Akiapola (H. wilsoni) and the Nihoa finch (Telespiza ultima) also show the (relatively rapid) beak radiation of endemic island birds from a common ancestor. The hope is, that in comparing the DNA base sequence of one species with another, a ‘molecular clock’ will prevail, i.e. that the number of base differences in any sequence from two birds will be directly related to the time since the two diverged. This will not be true if the length of DNA has some vital function but might work for lengths of DNA of no known function.

Liwi insects

Apapane

Amakihi

leaves

nectar

Nihoa finch

Akiapola’au

seeds

grubs

Founder species Maui parrotbill tool use

Figure 1.3 The adaptive radiation of beak forms in Hawaiian Honeycreepers. The following have kindly allowed permission to use their photographs or images within this book: Brent Cornell (BioNinja) https://ib.bioninja.com.au/standard-level/topic-5-evolution-and-biodi/52-naturalselection/adaptive-radiation.html

18

Biological Evolution: The Beginnings of the Story

Once again, the results with DNA sequencing of Darwin’s finches were short of convincing, but it became clear that the six ground finches were a natural closely knit group and that the warbler finches, comprising two species, had separated from all the others before the others had separated. In these studies, pursued by members of the Grants’ ‘school’, both nuclear and mitochondrial DNA were used, but in the 1990s a new molecular technique came into use. In the nuclear genome there are frequently considerable lengths of repeated short (2 or 3) base motifs (e.g. . . .CACACACA. . .) of no known function. This is called ‘microsatellite DNA’. Mutation consists of the addition or deletion of individual motifs (a single CA, for instance). So what characterises the microsatellite DNA of a bird is not one or more point mutations of single bases, but microsatellite length. Comparison of one microsatellite length from each bird would be of little help, but large numbers of these microsatellites are available from any individual and the whole set characterises that individual (a similar technique is used in so-called genetic fingerprinting for forensic purposes). In 1999, Petren, Grant and Grant were able to publish a phylogenetic tree of all species of Darwin’s finches, including the Cocos finch (Petren et al., 1999). In many ways the satellite tree corroborates Lack’s tree and the ordering into genera as listed above. But there are important differences of pattern and interpretation, to some extent foreshadowed in the results of DNA sequencing. One is the separation of the vegetarian finch, which branched off the main stock before the ground finches and the tree finches (including the woodpecker finches) separated from one another. A second inference is that the Cocos finch is not the first to diverge from the Galapagos finches – one of the warbler finches has that honour. Therefore, the ancestors of the Cocos finch almost certainly colonised Cocos Island from the Galapagos and not from the South American mainland. And lastly, and perhaps most importantly, the warbler finches themselves are only distantly related to one another. C. olivacea was the first to diverge from the whole stock, then the Cocos finch, and then the other warbler finch, C. fusca. All this happened before all the remaining species became distinct. This leads to a most important conclusion. The resemblances between the two warbler finch species are not indicators of closeness of relationship and must therefore be primitive for all Darwin’s finches. They therefore probably give us a good idea of the appearance of the first-ever Darwin’s finch and are a guide to our search for the bird species most closely related on the South American mainland. Lack did not know that there were two species of warbler finch (although he did know that ‘it’ was divisible into several possible subspecies). He correctly suggested early divergence from the main stock but thought that the ancestor of the whole group would be like one of the ground finches, with black male plumage and a crushing beak. In 2014 Skinner et al. explored epigenetic inheritance in five species of Darwin’s finch (see Figure 1.4). Epigenetic inheritance (more accurately transgenerational epigenetic inheritance) reflects the concept of heritable changes in which the physical structure of the DNA remains unchanged. The accepted model of course is that genetic mutation (established over time) generates the heritable phenotypic variation upon which natural selection acts. However, using erythrocyte DNA, this study explored

Classification and the Galapagos Finches

19

* Geospiza fortis 84 Geospiza fuliginosa 34

Geospiza magnirostris 161 Geospiza scandens 442

Geospiza conirostris Geospiza difficilis 606 52 Camarhynchus parvulus

Camarhynchus psittacula Camarhynchus pauper Camarhynchus pallida 1062 Platyspiza crassirostris 602

Certhidea fusca Pinarolozias inornata Certhidea olivacea

– Reference species Red – Epimutations (DMR) Blue – Genetic mutations (CNV)

*

Figure 1.4 Numbers of genetic and epigenetic mutations in relation to the phylogeny of the Galapagos finches (reproduced from Skinner M. K., Gurerrero-Bosagna, C., Muksitul Haque, M. et al. (2014) Epigenetics and the Evolution of Darwin’s Finches. Genome Biology and Evolution, 6, 1972–1989, by permission of Oxford University Press). (A black and white version of this figure will appear in some formats. For the colour version, please refer to the plate section.)

the possibility that epigenetic changes can accumulate producing the ‘clay’ upon which natural selection may introduce change. Blood samples were taken and analysed from five species of finch collected from Santa Cruz island in 2009. Genetic mutations were identified through use of gene duplication and deletion, while epigenetic variation was explored through differential DNA methylation analysis. The phylogenetic tree in Figure 1.4 follows Lack’s physical traits reinforced by Petren, Grant and Grant’s microsatellite data. Epigenetic mutations (epimutations) are shown numerically (in red) along with copy number variation (genetic mutations) in blue. There were generally more epigenetic mutations than genetic ones indicating that epimutations may be a major component in evolutionary change. There was also a significant correlation between the number of epimutations and phylogenetic

20

Biological Evolution: The Beginnings of the Story

distance – perhaps suggesting that epigenetic change accumulates over long periods of evolutionary time (1–3 million years). The relative importance of genetic and epigenetic inheritance remains a subject of considerable debate.

Darwin’s Finches and the Origin of Species A reconstructed pattern of evolution (as drawn by Lack for Darwin’s finches) is known as a phylogeny. Two historical processes must have contributed to that phylogeny. Any evolutionary change in time – in anatomy, behaviour, development or molecular constitution – is known by the term phyletic evolution. If phyletic evolution had not occurred in the history of life, then all living things would look like their first ancestors. The second evolutionary process (or ‘mode’ of evolution) is ‘speciation’, the splitting in time of one species into two or more. If no speciation had occurred in the history of life, there would be, assuming a single common ancestor, only one species of living thing on Earth! As we shall see later, there is still some dispute about the nature of speciation, particularly in animals. We have already seen that animal species are separated from one another by some sort of barrier, physical, behavioural or physiological, such that members of two different species do not normally interbreed with one another. This applies even in the case of two closely related (sibling) species occupying the same area, such as two species of Darwin’s finches living on the same small Galapagos Island. At some stage in their joint history a barrier to free interbreeding must have occurred. It has long been the claim of the evolutionary biologist Ernst Mayr that for that barrier to arise, splitting one species into two, there must have been an allopatric phase with the two, incipient species separated geographically. During the period that the two, incipient species are separated, differences in structure, behaviour, physiology, etc. evolve either in response to their different environments or by chance. If in the future the two species come to occupy the same habitat, the barrier to interbreeding will be enhanced by natural selection – Darwin’s great theory of the mechanism of evolution (see ‘The Galapagos Islands and Natural Selection’). Hybrids will either not be produced or will be at a competitive disadvantage to the pure-bred members of what are now separate species. Darwin’s finches have long been regarded as ideal exemplars of allopatric speciation (isolated by a physical barrier). A little thought will demonstrate that there are two ways in which part of a single species can be separated from the rest: 1. If the range of a species is divided by some geographical event, such as the opening of a seaway or the origin of a mountain chain, free communication between members of the species on either side will cease. This is a vicariance event. 2. If on the other hand a sample of the population crosses a preexisting barrier, either voluntarily or by chance (wind-blown, or the floating mats of vegetation referred to in the section ‘The Galapagos Islands and Darwin’s Finches: A Case Study’, for example), then this is a dispersal event. The islands were never connected to the South American mainland, from which they are far distant. They are also volcanic, so uninhabitable when first formed. But

Darwin’s Finches and the Origin of Species

21

dispersal from the mainland to the islands is certainly possible if one considers that the major Galapagos fauna are birds and reptiles (mammals are rare):  Birds of course can fly. Many of the coastal nesting birds (e.g. albatross) are strong, long-distance flyers. Others could have been aided by prevailing winds.  Reptiles are particularly hardy animals and could theoretically survive for many months on floating vegetation mats.  Mammals are much more vulnerable to water loss and lack of food and therefore less likely to survive long sea journeys. But the first dispersal event – South America to the Galapagos – does not explain the existence of 14 species on the islands. The finches are one of the best-known examples of adaptive radiation occurring due to allopatric speciation. Grant envisages: 1. The ancestral birds arriving on San Cristobal (Chatham). 2. Then after their adaptation to that island, a few birds ‘island-hopping’ to other islands to which they and their descendants became adapted. 3. The third and critical phase is when some island-hopping birds arrive on an island already inhabited by a finch population that has adapted to a particular way of life on that island. The new arrivals, if they survive, may have a different way of life and if breeding between the two stocks is inhibited, they will constitute sibling species descended from a common ancestor. The whole process depends on many islands, at least slightly different in their environmental demands, and the rarity of island-hopping events over many generations. Without a sufficient time-frame, the bird stocks will not be sufficiently different for an interbreeding barrier to be formed – they will still be the same species. When the two stocks are established, reinforcement occurs. Reinforcement is due to any factor that causes the two species not to interbreed, and to occupy different ecological niches. The whole process can be summed up as divergence in allopatry and reinforcement in sympatry. Sympatry is where two or more species overlap as they diverge. Thus, the theory of adaptive radiation by allopatric speciation implies that each species arose by differences in allopatry, presumably on a different island, and then reinforcement in (subsequent) sympatry. As a corroboration of this theory, evolutionary biologists point to the Cocos Island with its single finch species. There was no scope there for differentiation in allopatry after the ancestral birds had arrived. It is a characteristic of good science that however firmly established a grand theory is, scientists themselves will question it. Newton’s mechanics were eventually superseded by Einstein’s relativity. Therefore, can we question the scenario of an adaptive radiation of Darwin’s finches on the Galapagos Islands? The answer is yes, of course, but in this case that questioning will lead to a deeper understanding of the historical process, rather than destruction of the theory. (1) Until recently it was thought that no Galapagos island had ever been nearer to the South American mainland than those still existing. But in 1992, geologists discovered a series of seamounts (undersea extinct volcanic mountains) between

22

Biological Evolution: The Beginnings of the Story

the Galapagos and the mainland. Some of these underwater hills had cobbles indicating coastal erosion – they once protruded above the sea. So at least some island hopping between South America and the Galapagos might have occurred. (2) For the extreme allopatric model, inter-island flight should be a rare event, but occasional vagrant species are seen on islands where they are not regarded as resident, and in recent years (1982–1983) the large ground finch has coloured the small island of Daphne Major. (3) One case is known of the apparent beginnings of sympatric speciation, an island population dividing itself into two. The Island was Genovesa, small, remote and flat. The large cactus finch is resident on Genovesa and has been intensively studied by the Grants. They started their work there in 1978 after a severe drought in 1977. Male birds were heard to sing one of two distinct territorial songs, A or B. The nestling birds have one of two beak colours, yellow or pink. A males had 36% offspring with yellow beaks (and 64% with pink!), whereas B males had only 18% yellow. There was also a significant difference in bill length between adult A and B birds, correlated with different feeding habits. Type A birds fed on the flowers or hammered open the fruits of the prickly pear cactus, while B birds tore open the cactus pods searching for insects. There seems little probability that either A or B birds had arrived from elsewhere. The nearest population of the large cactus finch to that on Genovese is the only other one known, and is on the distant island of Espanola about 200 km away. There is no evidence of any other population ever having existed between the two. In subsequent years, the correlation between song type and beak length disappeared and there was no evidence that females of type A parentage mated for preference with type A males (assortative mating). Nevertheless, the division of the population in 1978 suggests the initial stage of sympatric speciation. (4) There have been many cases recorded of hybridisation between different species of Darwin’s finches, even those belonging to different genera. Over a period of 16 years, the Grants and their associates have conducted very detailed studies of the finches on the small crater island of Daphne Major. They discovered that the medium ground finches regularly hybridise at a low level with both the small ground finch and the small cactus finch, and that the hybrid offspring appear to be at no disadvantage, sometimes even flourishing more than pure-bred birds. This seems at odds with the idea that two sibling species in sympatry should evolve away from one another as the final stage of speciation. It also poses the question, too complex to deal with here, as to whether the six species of ground finches are in fact good species.

The Galapagos Islands and Natural Selection It was Charles Darwin’s great achievement not just to suggest that evolution had occurred (and give cogent reasons for accepting this) but also to propose a mechanism that could produce evolutionary change. His theory of natural selection proposed such

The Galapagos Islands and Natural Selection

23

a mechanism, answering in part the question, how can evolutionary change, phyletic evolution, occur, and how is it that living organisms adapt to changes in their environment? Natural selection is a theory of adaptive change. And yet, as we have seen, it was not generally accepted that the differences in size and shape of the beaks in Darwin’s finches were adaptively significant until the publication of Lack’s book in 1947. Darwin thought that natural selection, and thus adaptive change, was extremely slow and thus not open to observation. A study by Peter Boag and Peter Grant, published in 1981, showed that significant change can occur within a single generation. In 1977 the rains failed on the island of Daphne Major, resulting in very high mortality in the population of the medium ground finch, G. fortis, and no breeding. The birds that survived into 1978 were considerably larger than those that died (immediately showing the sex ratio, roughly equal before, of six males to one female). Seeds are the main food of G. fortis and in the drought year were in very short supply. This was particularly the case with the small seeds that were their staple. Only large birds with relatively deep beaks could open the seed cases and crush the seeds of the plant Tribulus. There was another plant with smaller seeds available, but this produced a sticky latex, gumming up the unfortunate bird. Thus, selection for the ability to eat large tough seeds resulted in a change in mean beak depth, but this would not result in evolutionary change, unless the difference were heritable. To show that this was the case the beak depths of later offspring were plotted on a graph against the mean beak depth of each offspring’s parents demonstrating that beak depth is indeed heritable. Adaptive beak depth can result from natural selection. A few years later the direction of selection was seen to reverse. In December 1982 the islands were hit by the effects of one of the most severe El Niño events of the twentieth century. The rains on Daphne Major continued to the end of summer and throughout the period G. fortis went into a breeding frenzy, each breeding pair, including some born in the same year, producing several broods. The population rose by some 400%. After the rains there came a population crash. But this time it was large birds (especially males) that were selected against. The reason is uncertain: there was in this case a surplus of small seeds relative to large, and it was suggested that, beak size or not, large birds simply needed more seeds, thus involving more searching to keep going. Also, it is said that the large beaks are less useful in the young before the beaks harden. The sensational changes in one generation resulting from both events on Daphne Major impressed evolutionary biologists, but the Grant team themselves noted that what they had described was a case of stabilising selection. The birds on Daphne Major could react rapidly to climate change, but over the years, body and beak size fluctuated about a mean. There was no evidence of selection producing phyletic evolution – sustained directional evolutionary change. For natural selection to produce such change, there has to be new genetic information on which selection can act. We will discuss the origin of such information (known as mutation) in Chapter 3.

24

Biological Evolution: The Beginnings of the Story

The Galapagos Islands therefore provide a powerful endorsement of Darwinian evolutionary theory along with an outstanding outdoor laboratory in which to study natural selection. The rate of evolution is rapid; in just under four million years there has emerged an amazing diversity of form arising from the animals and plants flying, floating, hitch-hiking or swimming there. Population sizes have become extraordinarily large in some cases, for instance in the unique marine iguana. Why should this be so? One possible answer is in the relative lack of predators on the island. Snakes are small and ineffectual against all but the smallest reptiles. The Galapagos hawk is limited by nesting sites, and among the invertebrates the large (20 cm) centipede Scolopendra is a formidable adversary for other invertebrates but it rarely troubles the islands’ vertebrate population. Large jungle predators from the South American mainland (mainly mammals such as the jaguar) just could not make the journey. Also, visitors to the island frequently note how tame all the animals are. Such ‘island tameness’ is typical of those creatures not subject to extensive predation pressure. It is suggested that a lack of predators coupled with smaller size and rapid reproduction (energy can now be diverted into reproductive success) has spurred the evolutionary trajectories of the Galapagos communities. Perhaps disconcertingly, a point was made of featuring data that might conflict with orthodoxy. Does the discovery of seamounts to the east of the present archipelago cast doubt on the accidental dispersal of ancestors to the Galapagos? Does inter-island flight occur too frequently for the allopatric scenario to be valid? Does hybridisation between apparent species refute their states as good species? Could incipient sympatric speciation, as seen in the large cactus finch on Genovesa, go to completion? I suspect that the answer to all these questions is ‘no’, but they must be asked. Progress in science is made by the recognition of data that appear to contradict established theories, by debate between theories and by continuous, ruthless questioning. Humans pollute – we can degrade the environment, but we also conserve. The last remaining giant tortoise on the island of Pinta, Chelonoidis abingdonii, was discovered in 1971 and relocated to the Charles Darwin Research Station on the island of Santa Cruz. Named ‘lonesome George’, this male Pinta giant tortoise was the very last of his kind. Attempts to locate other Pinta tortoises or mate him with females from closely related species all failed (although interestingly clutches of eggs were produced from such matings but none were viable). Sadly, Lonesome George died in 2012. The death of the ‘last of his kind’ was, of course, a blow but also a ‘wake-up call’ to conservationists globally. George’s death (at the relatively young age of 80 years or so) has reawakened conservation efforts both in the Galapagos and elsewhere resulting in a finding in 2015 that a closely related species, C. donfaustoi, had a 90% DNA match to that of George.

2

Reviewing the Evidence for Evolution

Evolutionary change has created biological lineages, all interrelated. And it is the role of the biologist both to describe the pattern of change and explain how our current biological diversity came to be. By observing present-day phenomena, we can see that:  animals and plants are grouped in a meaningful way in that they seem to form ‘natural groups’ (birds, fish, conifers, etc.),  there is a distinct heritability within living things (put simply, ‘like begets like’/ offspring resemble their parents),  an adaptation exists between an organism and its environment (living things seem to be well ‘fitted’ to the environment in which they live),  we observe ‘curious rocks’ that resemble living things that no longer exist (ammonites in mudstone or leaf impressions in coal) – we call these fossils. Darwin’s theory of evolution by natural selection remains the strongest argument in the explanation of all the above phenomena. It overturns Aristotle’s ‘fixed’ view of nature and replaces the notion of a preordained world. Modern science, from the sixteenth or seventeenth centuries onwards, replaces conjecture with evidence and dogma with reasoned argument. We live on a dynamic and rapidly changing planet, and we believe that living organisms also have changed – either in response to a fluctuating physical environment (and the opportunities or threats that affords) or to interactions between types of organism occupying those environments. But this point cannot be assumed or simply inferred; evidence is required. Over the years, several lines of evidence have been presented.

Homology and Comparative Anatomy By comparing animals and plants it is possible to infer relationships from their body structures. The feathers of birds, the milk of mammals, the organisation of sepals and petals in flowering plants all provide clues to their ‘connectedness’. But a closer examination reveals greater depth. The organisational plans (or ‘Bauplan’) of the various animal phyla reveal the potential for a common ancestor from which all members of that group are derived. These organisational plans (the type and arrangement of bones, common physiologies or behaviours) are the homologies on which this

26

Reviewing the Evidence for Evolution

section is based. The evidence they provide for evolution is indirect but overwhelming both in its logic and in its congruence with other lines of proof. If evolution was proposed by Darwin and Wallace to explain classification, then logically the existence of natural classification should not be cited as evidence for evolution. However, the pattern of evolution is important and fundamental to both classification and reconstructing evolution. Central to this notion is the concept of homology. The concept of homology predates that of evolution by hundreds, possibly thousands of years (being traced back to Aristotle in the fourth century BC). In 1555 Pierre Belon included a famous diagram (Figure 2.1) in his book of birds. He illustrated a human skeleton and that of a bird opposite one another, labelling in each what he considered to be corresponding bones. This correspondence, showing unity of plan, is a demonstration of homology. Put simply, homology refers to the basic similarity in structures observed in dissimilar species. The reason for homology is that these (dissimilar) species had a common ancestry and a basic structural pattern that has been preserved along the lineage. Let us take a standard example of homology, the pentadactyl (or five-fingered) limb of tetrapod (four-legged) vertebrates. In all but the very earliest tetrapods the limbs appear to be derived from a common pattern (Figure 2.2). There is a single bone that

Figure 2.1 A comparative anatomy of birds and Man from Pierre Belon’s (1555) book, The Nature of Birds.

Homology and Comparative Anatomy

27

articulates with a limb girdle; that bone then articulates with two parallel bones followed by the bones of the wrist or the ankle. Finally, hand and foot each bear five digits. This pattern then appears to have been modified to serve a range of functions in different tetrapods. For example, reduction to a single toe on each foot in the case of horses, or two toes (cloven hooves) in deer, antelopes and related animals. But the individual bones of the pentadactyl limb are homologous with one another (resemblance indicating common ancestry). The humerus of a horse is homologous with the humerus of a bat, although their functions relative to locomotion are very different (they are, of course, related to the same vertebrate ancestor). The whole forelimb of a horse is adapted for galloping, that of a bat for flight, but both appear to be derived from the common vertebrate plan of the pentadactyl limb. Homology therefore represents the state of affairs where a resemblance in body structure or morphology is the direct result of the organisms possessing a common ancestor and therefore acquiring similar structures (albeit often adapted now for different purposes). We can take the argument further by comparing the forelimb of a bat with that of a bird, the only other living vertebrate group capable of powered flight, and then ask is the forelimb of a bat homologous with that of a bird? The answer is ‘yes and no’! They are homologous as pentadactyl limbs, but not as wings, as should be clear from their structure. In birds the aerofoil surface of the wing is formed by flight feathers, and the bones of the hand, including the digits, are fused and reduced. In a bat the aerofoil

Figure 2.2 Homology as evidenced by the pentadactyl limb in vertebrates. (A black and white version of this figure will appear in some formats. For the colour version, please refer to the plate section.)

28

Reviewing the Evidence for Evolution

surface is formed by skin stretched over very elongated fingers and all five digits are present. If we now introduce a third animal for comparison, a house fly, also capable of powered flight, we can ask if the forewing of a house fly is a homologue of the wing of either a bird or a bat. The answer is a resounding ‘no’! And yet the fly’s wing subserves the same function as that of a bird and a bat – the means of powered flight. All three wings are analogous, that is they serve the same function. Analogy, a resemblance or superficial similarity, is common in nature as we see organisms arriving at the same ‘solution’ for the same biological problem (locomotion in water or seed dispersal in plant for instance). To return to the bird and the bat, their forelimbs are homologous as pentadactyl limbs, but analogous as wings. They have the same basic structure but are adapted in different ways for the same function. The great British comparative anatomist, Richard Owen, defined analogy and homology in 1843 as follows:

Analogue : A part or organ in one animal which has the same function as another part or organ in a different animal. Homologue : The same organ in different animals under every variety of form and function. Many biologists treat analogue and homologue as though they were mutually exclusive, but it is clear from Owen’s definitions that this was not his intention, nor is it proper usage today. A third term to describe non-homology is therefore necessary and was proposed by Ray Lankaster in 1870. It is homoplasy, a general similarity not due to common ancestry. The wing of a house fly is both analogous and homoplastic to that of a bird. There is no structural suggestion that both are modified from a common source. The most common cause of homoplasy is often convergence of DNA mutations. As a point of information, the different beak sizes of Darwin’s finches appear to be the result of change in one, or just a few, genes; thus, substantial adaptive radiation can be the result of relatively little genetic change. Of our three flight-powered animals the bird and the bat share many features as tetrapod vertebrates, including a forelimb which is a modified pentadactyl limb. But this character, to have a forelimb modified as a wing, is not a general feature of the lowest-ranking group to which they both belong (the Amniota). They are, however, much more closely related to one another than either is to the house fly. What we have here is a very small inclusive hierarchy of homologies. There are many more shared homologies between bat and bird, than between both of these and the fly. Thus, inferences of homology are the basis of classification. An order of precedence should now read: Data (i.e. ‘characters’) ➔ inferences of homology ➔ classification ➔ reconstruction of phylogeny (i.e. evolution).

In more modern terms, homologies used to define groups in classification are known as taxic homologies. But there is another category of homologies that can be used as evidence for evolution. It is transformational homology. As an example, we can take the three small bones that conduct sound from the ear drum to the inner ear in mammals.

Embryology

29

They form a chain, with the malleus (hammer) linked to the ear drum and followed by the incus (anvil), and the stapes (stirrup) which connects to the inner ear, where the organ of sound reception is situated. Comparative anatomy and embryology show that the malleus and anvil are homologues of the two bones (or cartilages) that form the jaw joint in all vertebrates except mammals. These are the quadrate at the back of the upper jaw and the articular for the lower jaw. In tetrapods the stapes is also an ear ossicle, but in fishes it is a bone (or cartilage) suspending the jaw joint from the braincase. So, the malleus and incus are transformational homologues of the articular and the quadrate, respectively. Such transformational homologues are inferred by the principle of connections proposed by the French comparative anatomist Étienne Geoffroy St-Hilaire (1772–1844). Homologues in two animals might look very different and serve very different functions, but they can be recognised by their connection to surrounding structures and their general topology within the respective organisms. We shall see below that transformational homology can be combined with the fossil record to provide powerful evidence for evolution. For completeness we should note another type of homology, which was very much the focus of attention at the turn of the eighteenth/nineteenth century. It concerns homology within a single organism and was named ‘serial homology’ by Owen. In a centipede all the many body segments, after the specialised head region, are remarkably similar and are clearly built on a common plan – each is a serial homologue of all the others. The same serial homology applies, but rather less so, to the body segments of a vertebrate. Each segment possessing a vertebra, a pair of ribs, a pair of spinal nerves and segmented muscles. But what has happened at the head end? It was suggested by the great German writer Johann Wolfgang von Goethe (1749–1832) that the vertebrate skull consisted of (probably four) modified vertebrae. This was accepted by many anatomists including Owen but is not generally believed today. Another theory of Goethe’s is however accepted. In 1907 he suggested that for the parts of a flower, each of the sepals (collectively the calyx), petals (corolla), stamens and pistils, could be regarded as a modified leaf. Such homologies are now known as iterative homologies as they are not necessarily arranged in serial order.

Embryology As with the study of comparative anatomy, a comparative study of animal embryos reveals similarities in form that puzzled early biologists. It was Charles Darwin who reconciled these strange and yet similar developmental processes under the unifying theme of biological evolution. Once again, the argument is indirect but compelling. During the eighteenth century, and in the first half of the nineteenth, the majority opinion among naturalists was not that the natural order of organisms was a tree-like inclusive hierarchy, but rather a linear series, a ladder or ‘scala naturae’. This ‘Ladder of Nature’ or ‘Great Chain of Being’ was pictured with Man near the top and going down through ‘higher and lower’ animals to plants at the bottom along with inanimate objects such as rocks and pebbles (Figure 2.3).

30

Reviewing the Evidence for Evolution

Gods Angels Man Woman Animals Plants Minerals Figure 2.3 Aristotle’s ‘Great Chain of Being’ or ‘Scala’.

In human development, the growing embryo was thought to climb the scala naturae individually (or at least the upper vertebrate part of it). Thus, the individual developed through stages corresponding to fish, amphibian, reptile and mammal before birth. The person who demonstrated that this was false was Karl Ernst von Baer, who in 1828 showed that the pattern of embryo development was not through a linear series, but one of divergence. Each embryo of a given species instead of passing through the stages of other animals departs more and more from them. Fundamentally therefore, the embryo of a higher animal is never like (the adult of ) a lower animal, but only like its embryo.

Here then there appears another irregular divergent hierarchy that could be used as a basis for classification. According to von Baer, the early embryos of two different species share many of their characters but diverge by acquiring more and more specialised characters as they develop. High-ranking characters develop first, successively lower ones later. This is not universally true: a hen’s egg has many special features (adapting it to be free of standing water) that develop before features characterising all tetrapods, and other embryos may have features that ensure their survival, but in general terms von Baer’s Laws give a truer picture of ontogeny (individual development) than the scala naturae. A rather appealing (though incorrect) hypothesis concerning animal embryos and evolution is Haeckel’s ‘biogenetic law’, sometimes referred to as his ‘Theory of Recapitulation’. The similarity between ‘higher’ vertebrate embryos (see Figure 2.4) is explained as the animal somehow recapitulating its evolutionary history. Ernst Haeckel, a German evolutionist and contemporary of Charles Darwin, noted the resemblance between animal embryos and the adult form of its ancestors, while German Naturphilosophie at that time saw deep bonds between individual

Embryology

31

Figure 2.4 Comparison of vertebrate embryos noting their similarity, (Pig, Cow, Rabbit, Human).

32

Reviewing the Evidence for Evolution

development and the organisms’ evolutionary development – as exemplified in Haeckel’s phrase ‘ontogeny recapitulates phylogeny’. Haeckel’s Biogenetic Law (Haeckel, 1866) is one of several recapitulation hypotheses. The resemblance between embryos, he argues, is that the animals are ‘replaying’ their phylogenetic history. Stanley Hall in the United States took this idea into the realm of developmental psychology arguing that children go through early stages of development paralleling their evolutionary development, and biologists Ernst von Baer and Louis Agassiz were strong advocates of this type of biological progress. It is not clear whether Charles Darwin was influenced by this type of German metaphysical thinking, but the general opinion is that he was rather wary of ascribing a distinct biological progress to his findings. In Darwin’s eyes progress came about naturally and organically through the process of selection. Ernst Haeckel proposed his biogenetic law after reading Darwin’s Origin of Species. However, where Haeckel attributed embryonic similarity to extant species resembling the adult forms of their ancestors, Darwin suggested the embryos of current forms look similar to those of their ancestors because they all shared a common ancestor. Haeckel’s Theory of Recapitulation has largely gone out of fashion since it became clear that embryos do not resemble the adult forms of their forebears, rather they resemble their ancestor’s embryos. Similarly, there is no fixed relationship between embryo and adult form; not all characters are seen in the embryo, and in some groups, like the plants, all traces of their ancestry are lost during development. But although recapitulation does not provide an explanation, there are legitimate grounds for inferring that the resemblance we see between embryos is the result of a shared ancestry and therefore provides strong evidence for evolution along the lines of embryological homologies.

Vestigial Organs One transformational homology was cited by Darwin as evidence of evolution, that of vestigial organs. The hip girdle and hind limb of whales are entirely embedded within the body yet are clearly homologues of the functional back legs of other mammals (Figure 2.5). Similarly, the baleen or ‘whalebone’ whales, such as the blue whale, lack teeth. The baleen whales feed by filtering sea water to retain plankton through a series of long narrow plates of baleen, made of keratin, the stuff of hair and fingernails. Yet the embryos erupt apparently functionless teeth which are soon lost. Primitive snakes such as the boa constrictor also have reduced hind limbs (although they have lost their forelimbs altogether). The hind limbs appear externally merely as spurs, either side of the vent. Vestigial traits are a feature of a species that once had an adaptive purpose in the ancestor of that group but have now either lost their use, as in the case of the human coccyx, or have been coopted for a new use such as the early vertebrate jaw bones now found in the inner ear. Structures such as tiny hind limbs in snakes, nonfunctional eyes in cave-dwelling fish and air sacs in the limb bones of flightless birds are vestiges of a former, ancestral, lifestyle.

Vestigial Organs

33

Figure 2.5 Whales possess limb bones similar to that of hooved mammals.

An example of what Darwin confusingly called ‘rudimentary organs’ is the useless wings of some beetles. In beetles the forewings, present in most insects, are modified as hard wing cases not used in flight but opened down the midline so that the hind wings can operate. But in some species of so-called ground beetles the wing cases are, in Darwin’s phrase, ‘soldered together’ so that the hind wings, still present, are useless for flight. In this case there seems no possible explanation other than that the ground beetles are descended from flighted ancestors. Another extraordinary case involves the recurrent laryngeal nerve of mammals. This nerve is a branch of the vagus nerve. The vagus emerges from the brain near the back of the skull. The recurrent laryngeal branch then runs along the neck, loops around the dorsal aorta (the large blood vessel that carries blood from the heart for distribution round the whole body except for the head and neck) and then redoubles back towards the skull, providing the nerve supply to the larynx. If the human recurrent laryngeal ran straight from the skull region to the larynx it would be centimetres shorter – the recurrent laryngeal of the giraffe would be several metres shorter! The reason for this apparently wildly inefficient piece of wiring is clear if one looks at the situation in fish, who have no neck. The dorsal aorta of mammals is the homologue of the fourth (from the front) of a series of blood vessels supplying the gills. Numbers 1, 2 and 5 have disappeared in the adult mammal: number 4 is the dorsal aorta and number 6 the pulmonary artery taking blood to the lungs. In the embryo mammal numbers 4 and 6 are connected and the recurrent laryngeal loops behind the connection (the ductus arteriosus). The ductus degenerates in the new-born mammal, but the nerve continues to loop behind the ligament that is its remnant and thus around the aorta. In fish the nerve goes straight to its destination. In each of these cases of vestigial organs one is compelled to see the anomaly as a remnant of an ancestral condition – phyletic evolution (the divergent development of two or more daughter species from a single parent species) has occurred. But a note of caution is necessary. To take a previous example, many cave-dwelling organisms, such as shrimps, fish and the blind cave salamander Proteus, have degenerate eyes or none at all. But all have close open-air relatives with eyes normal for their group.

34

Reviewing the Evidence for Evolution

Proteus (the Olm), which occurs in caves near the eastern Adriatic coast, is also neotenous, that is, it lives in a permanent larval condition, with feathery external gills and limbs with reduced digits – three at the front, two at the back. It thus seems probable that the eyes are vestigial – they develop so far and then degenerate. But in other cases, it may be more difficult to decide which way to read the series. All such cases are evidence for evolution, but one needs to know which way to read the transformational homology. In answer to ‘special creationists’, who claim that a complex coordinated organ such as the human eye could never have evolved by natural selection, the evolutionary biologist Richard Dawkins has presented a compelling scenario, with each stage better adapted for vision than the one before. Many protozoans (single-celled organisms) have an eye-spot of one or a few light receptors, and so do some animals, with in this case several light-sensitive cells. In some cases these cells are partially shielded, so that not only is the creature aware of light and darkness but, by movement of the shadow, the direction of the light. Next, we move to the light cells situated in a hollow cup and omnidirectional detection. A deep cup with a small aperture gives a poor, dim and fuzzy, pinhole camera effect. Decrease the size of the pinhole and the image is sharper but dimmer; increase it and the brightness improves but the image worsens. Oddly the living marine mollusc Nautilus has only reached this stage, while the related octopuses and squids have evolved the solution, a lens; though their eye is very different in construction and embryology from the vertebrate eye. Animal ‘vestiges’ are still used by creationist thinkers to try and disprove evolution. Why, they argue, has an animal (or plant) evolved a useless feature? Alternatively, some argue that vestigial features do have a function (possibly one not discovered yet) and that the biologist is hopelessly trying to find an evolutionary antecedent. In other words, a trait cannot be vestigial if it still has a function. But what is missing in these arguments is that evolutionary theory does not say that a vestigial organ has no function – the point is ‘it no longer performs the function for which it evolved’ (Coyne, 2009). The point of this digression is to pose the question, how, when faced with a poorly developed structure in an animal relative to its near relations, do you tell whether it is vestigial or not yet fully developed (‘rudimentary’!)? The answer again is to look at the classification. A blind, cave-dwelling creature with poorly developed eyes (or no eyes!) is likely to show a degenerate, or vestigial, character, if open-air forms, who are also members of the group to which it belongs, have fully developed eyes. Another possible explanation of a strange vestigial feature is the ‘throwback’, the reappearance of a character once lost. It is said that Julius Caesar for instance once rode a horse with feet ‘almost human’. And both sperm whales and humpback whales may be found with ‘projecting rudimentary hind limbs’. The reappearance of a previously functional characteristic is known as an atavism. Vestigial traits occur naturally within a population whereas atavisms are often spontaneous (but still they shed light on evolutionary history). The existence of hen’s teeth is perhaps a myth but in 1980 it was demonstrated that grafted mouse mesenchyme tissue could induce dentine formation in chick enamel suggesting that genes might be conserved unaltered for millions of years yet remain inactive.

The Fossil Record

35

The Fossil Record The usual assumption is that the fossil record is the evidence for evolution. It certainly represents evidence of a change in the totality of living things (the biota) over time. But to make the case most forcefully, one needs to produce a logical argument in several steps. In this section the assumptions will be itemised and the argument (that the fossil record is evidence for evolution) presented in full: 1. that fossils are evidence of once-living organisms and not some parallel phenomenon; 2. that fossils in their stratigraphic setting (i.e. their sequence in the rock strata) constitute a historical record; 3. that extinction is a real phenomenon; 4. that there is progression in the fossil record, i.e. that the record shows a succession of biotas through geological time, such that individual species in the record are clearly related to those that come before and after them, without being identical; 5. that, despite known major extinction events, the fossil record is not interrupted by any catastrophic event that is known to have wiped out all existing life; and, furthermore, that the first appearance of any fossil group (from species upwards), and their disappearance from the record, do not always coincide with known major extinction events; and therefore 6. that any apparent progression (which need not imply improvement, just change) is more plausibly interpreted as phyletic evolution than as ‘catastrophism’ (i.e. the extinction of the whole biota to be replaced by a newly created one). The argument regarding the use of fossils is explored further below. The same number order is employed. 1. The term ‘fossil’ is used to label any specimen preserved in a rock stratum, or other ancient material such as amber, that constitutes evidence of a former living thing or things. Some fossils seem to us today to be easy to interpret. Fossil mollusc shells, bivalves or snail-like, are so like living forms, and sometimes of the same species, that their interpretation seems obvious. But two things must be borne in mind. Firstly, as was written by a philosopher of science, D. B. Kitts, in 1974: ‘fossils by themselves tell us nothing; not even that they are fossils’ (Kitts, 1974). Directly or indirectly, fossils are always interpreted by comparison with living organisms. Secondly, fossils have not always been interpreted as the remains of once-living organisms. During the seventeenth and eighteenth centuries there was a vigorous debate between those who advocated the modern view and those who believed that fossils represented a parallel creation, never alive but simulating living things. In the 1660s a Dane (Nils Stenson) offered two reasons for interpreting fossils as once living. He looked at the common ‘tongue stones’ (petrified shark teeth) of northern Mediterranean strata and noted that they were closely similar, in both external form and interior construction, to the teeth in living sharks. His second criterion was that the fossil teeth embedded in rock had impressed their form on the surrounding rock

36

Reviewing the Evidence for Evolution

sediment. They were therefore formed before being buried in the sediment. They had not ‘grown’ by some means within the sediment itself. But those who doubted the nature of fossils had their reasons. Finding fossils of creatures no longer alive implied extinction, which cast a slur on God’s creativity – He would not create organisms only to scrap them. They were also disturbed at finding fossils of marine shells high in mountains – now known to be due to tectonic movements within the earth. 2. Historical geology, as taught to all students of the subject, depends heavily on two principles, the principle of superposition and the principle of correlation. The first is delightfully simple, but all important. In an exposed sequence (a ‘section’) of sedimentary rocks (those formed by deposition of preexisting particles), such as sandstone or mudstone, unless the section has been radically disturbed, the oldest will be at the bottom. The second principle is used to judge the relative ages of strata throughout the world. If a stratum in one part of the world has a fossil biota the same as or close to that of a stratum elsewhere, then the two strata are of approximately the same age. The converse does not apply: two strata from different localities with very different biotas are not necessarily of significantly different ages: they could be of the same age but represent very different environments – one marine and one freshwater for instance. Since the middle of the eighteenth century, geologists have painstakingly been reconstructing the sequence and thus relative ages of strata throughout the world, and the work continues to this day, so that we now have considerable knowledge of the relative ages of sediments worldwide and thus of the sequence in the fossil record of any group of organisms in which we are interested, together with a hierarchical classification of stratigraphic time, dividing the whole Earth history into aeons, divided into eras, divided into periods, divided into epochs, divided into stages. And so, fossils found at different levels in the sequence can be placed in time relative to one another. Some of those who choose not to accept that evolution has occurred deny the historical nature of sedimentary dating. But those people now have to confront the acceptability of modern physics as well as all historical geology and systematic biology. During the twentieth century a number of techniques were developed for radioactive dating, using measurement of the rates of decay of radioactive elements to give the ages in years of rocks and directly or indirectly of fossils. 3. As we saw above, pious scientists in the seventeenth and eighteenth centuries had great difficulty in accepting the reality of extinction. To them fossils were either a parallel creation, probably never alive, or, alternatively, creatures known only as fossils would one day be found alive as the world became more fully explored. Most fossils then known were of marine animals, and it was easiest to believe that someday they would turn up from the depths. But at the end of the eighteenth and the beginning of the nineteenth century, large fossil mammals, mammoths, ground sloths and others were discovered in both the Old and the New World. It was very unlikely that such conspicuous land animals had been overlooked as the Earth became better known. Many of these had been described as fossils by Georges

The Fossil Record

37

Cuvier (1769–1832), the founder of systematic vertebrate palaeontology, and it was he who convinced the scientific world of the reality of extinction. 4. It is also Cuvier, with his fellow French geologist Alexandre Brogniart, who share the credit with an Englishman, William Smith (1769–1839), for establishing the principle of correlation (see above) for the relative dating of fossils. Thus, in a rock section, if each stratum has its own distinctive suite of fossils, it can be assumed that those in a lower stratum had become extinct before the preservation of those in the stratum above. A good example is the series of ammonite zones to be found at the beginning of the Jurassic period in the rocks of the north Somerset and south Dorset coast. But unless the biota of the world has been steadily diminishing since the earliest fossil record, all the new fossils characterising a zone must have come into existence at or before the beginning of that zone. There has been change (or ‘progress’) in the fossil record throughout, with one biota replacing the immediately previous one. 5 and 6. Two obvious explanations suggest themselves for the progression. Cuvier and other early nineteenth century palaeontologists believed that the history of the Earth was to be a series of catastrophes that wiped out all living things and that after each catastrophe the biota was re-created with slightly different animals that survived until they themselves were wiped out. The other explanation was that the younger biota arose from the older by phyletic evolution and speciation. Two arguments favour evolution; firstly, there is no probable explanation of the whole series of creation events that would be needed, nor is there any evidence of such a creative event. Secondly there has been a small number of major extinction events in the history of the Earth (of which the most famous, but not the most catastrophic, was that at the end of the Cretaceous Period, ca. 65 million years ago – extinguishing all the remaining dinosaurs – but see below!). But there is no evidence of any extinction event that wiped out all the biota. Furthermore, there is what is known as a ‘background rate’ of extinction. Fossil species become extinct (and new species appear) regularly in the record, not just coinciding with major extinctions. This second point was made forcefully by the geologist Charles Lyall (1797–1875) in the third volume of his Principles of Geology (1833). He set out to divide the era from the Cretaceous extinction to the present day into four epochs. But rather than using the classical correlation methods of Smith and Cuvier, Lyall used a statistical technique. Using a vast data base of living and fossil mollusc shells, he defined each epoch in terms of the percentage of fossil shells found that were still known alive today – Eocene ca. 3%; Miocene ca. 20%; Older Pliocene 33–50%; Newer Pliocene ca. 90%. Thus, while many new species must have evolved since the beginning of the Eocene, there could have been no time at which an extinction event wiped out the mollusc fauna, contrary to Cuvier and Briognart’s study of the same era. Thus the ‘catastrophism’ of Cuvier and others can be rejected, and change from one stratum to another in the fossil record is evidence of evolution. One category of fossils has always been of interest to evolutionary palaeontologists. It consists of extinct organisms that span the difference between what are now

38

Reviewing the Evidence for Evolution

distinct taxonomic groups. The most famous of such fossils is Archaeopteryx from the Upper Jurassic of Southern Germany. Some half a dozen skeletons are known, preserved in the fine lithographic limestone of Solnhofen region. Archaeopteryx is usually presented as a ‘missing link’ between reptiles and birds but should more correctly be thought of as a bird that has retained primitive reptilian features. The most important bird feature is the presence of beautifully preserved feathers; not just contour feathers surrounding the body, which may have been present in some dinosaurs found in China, but true flight feathers on the wings, so diagnosed because they are asymmetrical with an off-central shaft and arranged as primaries and secondaries as in living birds. Otherwise Archaeopteryx is very much like small carnivorous dinosaurs. It has teeth in its jaws (seen also in later fossil birds); there is no massive sternum for the origin of flight muscles, and the long bones are solid rather than hollow as in birds. Its ‘hands’ retain three long flexible fingers with claws, and there is a long tail of vertebrae but fringed with feathers. However, the legs are more like those of living birds. To sum up, Archaeopteryx is not only a feathered dinosaur, but also a ‘stem-group bird’, demonstrating that birds are in fact a subgroup of dinosaurs.

Fossils and Phylogeny Having demonstrated in general terms that the fossil record embodies evidence for evolution, two examples of individual evolutionary series are now presented. The first involves the evolution of an anatomical structure, using the concept of transformational homology, and the second an evolutionary series of animals. For the example of transformational homology, we can take the three ear ossicles of mammals (see Figure 2.6) and trace their history in the fossil record. There is, first, evidence of the homology of the quadrate and the incus, and of the articular and the malleus from the embryology of mammals. In the foetus of primitive placental mammals, such as the hedgehog, the malleus can be seen developing in continuity with Meckel’s cartilage; the cartilage then forms the axis of the lower jaw in vertebrates. Only later does the malleus become separate. The incus is then loosely attached to the malleus, and the stapes to the incus. There is also a ring-shaped bone, the tympanic, that holds the mammalian ear drum, into which the malleus inserts. Marsupial (pouched) mammals, such as the kangaroo, are, in most structures, more primitive than placentals such as the hedgehog. The baby kangaroo is also borne at a very early stage of foetal development, so that when the newly born ‘Joey’ suckles its mother, it is the primitive reptilian jaw joint that is used. The young kangaroo swaps from the primitive quadrate-articular articulation to the diagnostic mammalian one in the early days of its life. Turning to the fossil record, the first creatures showing any diagnostic mammalian features occur in the Carboniferous period, some 315 million years ago. The first mammals with all the diagnostic characters appear in the Upper Triassic just over 200 million years ago. Between these there is a rich record of what are technically known as stem-group mammals. These do not constitute a single lineage from the

Fossils and Phylogeny

Mammal

squamosal

inner ear

stapes incus (quadrate)

eardrum malleus (articular)

dentary

squamosal quadrate

Reptile

39

articular eardrum

inner ear

stapes

dentary Figure 2.6 Evolution of the auditory ossicles (ear bones) of mammals from the jaw. (A black and white version of this figure will appear in some formats. For the colour version, please refer to the plate section.)

most primitive stem-group mammal to the earliest true mammal, but a rich bush-like pattern. Nevertheless, it is possible to see every stage in the evolution of the middle ear of mammals from the jaw region of early tetrapods. The quadrate (incus) is reduced in size and loosens in the skull. The malleus (articular) is similarly reduced. A jaw bone called the angular is converted into the tympanic, while other lower jaw bones are lost leaving only the dentary, the sole lower jaw bone of mammals, which grows back to form the jaw joint with the top of the skull (Figure 2.6). Note how in the reptile the quadrate and the articular still form part of the jaw (again strong evidence for homology and for similar ancestry). Once evolution was accepted among at least some palaeontologists in the late nineteenth century, it was assumed that by searching the fossil record, one could find direct evidence of phyletic evolution in the form of series of fossils through time representing ancestor–descendant sequences. The simple idea of the scala naturae turned into evolutionary history. Probably the most famous example of this was the evolution of the horse. About 55 million years ago the earliest known member of the horse family (Equidae) flourished in the early Eocene epoch. Usually known as ‘Eohippus’ it is more correctly known as Hyracotherium. There are many species, the smallest of which is about the size of a domestic cat. Today there is only a single horse genus, Equus, that includes domestic and wild horses, asses and zebras. It was natural that when a number of apparently intermediate forms were found that the whole collection should be

40

Reviewing the Evidence for Evolution

Equus

Pliohippus

4 mya Parahippus Miohippus

55 mya

Hyracotherium

Figure 2.7 Evolution of horse limbs.

arranged in a linear series, not only to demonstrate the evolution of modern horse characters but as an actual ancestor–descendant sequence. But, in fact, it is now known that the history of the horse family shows a much more complex pattern. There was not just a single evolutionary line from ‘Eohippus’ to Equus but a branching ‘bush’ with, notably, a great radiation of browsing horses with lowcrowned molar teeth, coexisting with early grazing horses with high-crowned molars with self-sharpening vertical plates of dentine, enamel and cementum, the present condition. One of the main differences between ‘Eohippus’ and Equus is that the former has four toes on the forelimb and three on the hind, while Equus has one on each. The reduction in toes is seen in Figure 2.7 along with the (predicted) transitional forms. Nevertheless, there must have been a line of descent from Hyracotherium or something closely related to it to the living genus, and if one could demonstrate the acquisition of the characters of Equus in sequence through time, it would provide powerful evidence of phyletic evolution in the horse family. One can never be sure that one has an ancestral sequence in the fossil record, but one can, given the evidence, reconstruct the timing of those ancestors by classifying the actual fossils. As a very simplified example, Figure 2.8 represents the classification of just seven extant (or living) horse genera, using the states of just four characters: body size, number of toes on the fore-foot, the presence of high-crowned teeth and the presence of a postorbital bar. Now if the pattern of classification (a cladogram) is regarded as a tree, in other words the pattern of evolution, each branch-point or node represents the ancestor of all

Biogeography

41

E. quagga E. burchelli E. grevyi E. zebra E. africanus E. hemionus E. pzewalskii (E. caballus) 3.9

mya

0

Figure 2.8 Cladogram showing the seven, living species of Equidae (genus Equus).

the branches that descend from it. Using complete mitochondrial sequences, Vilstrup et al. (2013) explored modern equids showing relationships, for instance, that zebras are “monophyletic within the genus”, that is they share a common ancestor.

Biogeography Biogeography (the geographic distribution of animals and plants) is not random but a result of both environmental and evolutionary factors. Yet the evidence from geographical distribution was instrumental in the work of Charles Darwin. He provided examples from Australia, South America and South Africa, all of which provide similar environments yet dissimilar flora and fauna. Darwin also drew attention to animals and plants on islands. They generally have a flora and fauna less rich than the mainland nearby and several species endemic (that is unique) to that location. A major question for nineteenth century biologists though was how these animals and plants managed to get dispersed at all. Darwin himself spent many years conducting experiments on how snails and seeds could withstand saltwater and may be carried by floating matter or transported on the feet of birds. An alternative hypothesis was proposed by Darwin’s close friend, the botanist Joseph Hooker. He explained that plant and animal distribution might well have been affected by land bridges in the past, now covered by rising sea levels. A third view was that different biotas might have been produced by a special creation at each location! But this seemed hardly plausible to Darwin and, of course, did not explain the relatedness seen between species on different islands (for instance the finches discussed in Chapter 1). When discussing fossil evidence for evolution earlier in this chapter, a series of observations were listed that led to a strong evidence base. This process is repeated and presents the argument for biogeography as a major indicator of evolutionary change: 1. The pattern of distribution of plants and animals cannot be explained by environmental factors alone: their history must be invoked. 2. Comparison of geographically distant but similar environments makes it probable that each species developed separately in the environment to which it was most highly adapted.

42

Reviewing the Evidence for Evolution

3. All other things being equal, the resemblance between any two floras or faunas is inversely related to the width and/or effectiveness of the barrier between them. 4. Geographical regions can be arranged as a hierarchical classification (regions within regions) based on classification of the different organisms that inhabit them, and that regional classification can be corroborated by geological events. The French botanist Augustine de Candolle was one of the first to attempt to put biogeography on a scientific footing. In 1820 he pointed out the distinction between what he called ‘stations’ and ‘habitations’ of plants. Thus, the station of a plant is the immediate type of environment in which it flourishes – sundew in acid bogs, water crowfoot in ponds and streams, mahogany in tropical rain forests, marram grass in coastal sand dunes, and so on. But habitations require some other sort of explanation – why does poison ivy occur in all sorts of environments in temperate North America but not in Europe, or magnolia trees in south-eastern USA and Central America, as well as in south-east Asia, but not at similar latitudes in between? Or for that matter, why do some plants, such as the plantain occur in suitable habitats anywhere in the world? De Candolle then divided the world into first 20, then 40 regions of endemism, i.e. major regions to which part of their flora was exclusively confined. He was followed by zoologists defining land regions. In 1858 Philip Sclater (a prominent lawyer and zoologist) suggested six land areas, based principally on birds, and this pattern was later adopted by Alfred Russel Wallace in his (1876) Geographical Distribution of Animals. In most cases there is an obvious barrier between these zoogeographical regions. The Atlantic Ocean separates the two New-World regions from the Palaearctic and African regions, which are separated by the Sahara and the deserts of the northern Arabian Peninsula. The Oriental region is outlined in the north by the Himalayas. The separation though of the Oriental and Australian regions, with very different mammals and birds, was a mystery to Wallace and others and has only been explained in the mid-twentieth century with the theory of plate tectonics. But why should faunal regions be evidence of past history and thus for evolution? Perhaps there are important differences between the forests of South America and Africa, with jaguars and New-World monkeys designed to fit the former and leopards and Old-World monkeys the latter? The history of Australia gives the lie to this precept. Humans have introduced a whole species of placental mammals – mice, rabbits, pigs, buffalo and even camels – into the island continent, and they have flourished often at the expense of the native marsupials. As Darwin puts it in the Origin of Species: No country can be named in which all the native inhabitants are now so perfectly adapted to each other and to the physical conditions under which they live, that none of them could anyhow be improved; for in all countries the natives have been so far conquered by naturalised productions, that they have allowed foreigners to take possession of the land. The fact that the distribution of animals and plants demands a historical explanation is emphasised by the comparison of the faunas of islands with the nearest mainland. The Galapagos Islands off the west coast of South America and the Cape Verde islands off the west coast of Africa are very similar. Both archipelagos are tropical

Observational and Experimental Evidence

43

volcanic groups and have never been connected to the mainland. They have a similar climate. Both lack amphibians and naturally occurring mammals, which cannot easily disperse over oceans. But in each case the fauna is obviously derived from that of the nearest continent. The Cape Verdes have a bird fauna, notably with several species of kingfisher, derived from that of West Africa, while that of the Galapagos (including Darwin’s finches) is derived from that of South America, as are other elements of the fauna such as the land and marine iguana lizards. This leads us to another comparison. We can compare the biota of distant oceanic islands, with reference to the nearest continental mainland (as with the Galapagos and the Cape Verdes) and close continental islands such as the British Isles. The British fauna is principally an impoverished version of that of continental Europe; unique mammals are merely endemic subspecies while in the Galapagos Darwin’s finches are an endemic subfamily. This comparison suggests another hierarchy. Britain and Western Europe could be compared to two geographical ‘species’ within the same ‘genus’: the Galapagos fauna is sufficiently different from that of South America to give the Galapagos ‘genus’ status (with the individual islands as ‘species’?). On the world scale of land animals, the Sclater/Wallace regions would rank as the highest groups. In an essay published in1855 (three years before his joint presentation with Darwin of the theory of natural selection) Wallace said: Large groups, such as classes and orders, are generally spread over the whole earth, while smaller ones, such as families and genera, are frequently confined to one portion, often to a very limited district. He then goes on to claim that the degree of difference between the fauna on either side of a natural barrier is proportional to the effectiveness of that barrier and to its age. Wallace was obviously reaching for the idea of a biogeographical classification, but such a classification could only be realistically undertaken with two ingredients that only became available late in the twentieth century. These were the theory of plate tectonics and the rigorous method of classification known as cladistics. Plate tectonics, originally known as ‘continental drift’, is the theory that the Earth’s crust is divided into a number of rigid plates, including ocean floor and the emergent continents. Plates grow at the volcanic mid-ocean ridges and are destroyed in subduction zones as they slide under the continental masses of other plates. As a result, over geological time, the continents have moved relative to one another. Thus, in our area classification, adjoining continents have split apart dividing originally connected animal or plant distributions – a vicariance event (splitting of biota into discontinuous parts). Subsequent evidence of such an event is that many unrelated groups of organisms appear to have been divided into separate species at the same time by that event.

Observational and Experimental Evidence Evolutionary adaptation is seen in the house sparrow (Passer domesticus). Native to Europe and the Middle East, this species has been introduced throughout the world.

44

Reviewing the Evidence for Evolution

From an initial breeding population in New York in 1852 the birds spread rapidly throughout North America undergoing evolutionary change in both size and colour. Northern populations are darker and larger than those in the southwest, while an adaptation to hot humid climates in the south has resulted in lower metabolism and greater insulation. This example of slow, small-scale evolutionary change is well documented by Johnston and Selander from the 1960s onwards, while the consequence of increasing size (increased wing loading) is countered by directional selection on wing area (Monahan, 2008). There are several such examples of evolution-in-action, but examples of evolution can also be produced experimentally. Changes in the fruit fly (Drosophila melanogaster) have been induced through selective breeding and subjecting flies to adverse environments such as low oxygen conditions. The short generation time of bacteria has also been employed to grow on substrates or to ‘swarm’ much faster than normal. Long-running experimental evolutionary experiments include Garland’s (1998) selection of mice for running ability (65 generations), Rose’s (1984) selection of multiple fruit fly traits indicating fitness (over 200 populations demonstrating evolutionary radiation) and the longest running of these (1988 to the present day), Lenski’s Escherichia coIi experiment. From 1988 Richard Lenski has been tracking bacterial changes in 12 initially identical bacterial populations. In one investigation bacteria were encouraged to grow on a citrate substrate rather than glucose. Eventually a spontaneous citrate metabolising mutation (cit+) arose after around 31 000 generations (see Michigan state University website for an overview http://myxo.css.msu .edu/ecoli/overview.html).

3

Genetic Variation within Populations

One of the most arresting aspects of the living world is its diversity. No two individuals of a sexually reproducing population are the same; neither are two populations of the same species, two different species or any two of the higher taxa. The biological diversity of individuals within a species is the subject of this chapter. Historically, the description of individual variation has been of interest to naturalists and farmers for centuries. A study of the origins of individual variation however is a recent phenomenon, contingent on our knowledge of inheritance and heredity.

Inheritance and Variation Ladybirds, or ladybugs, are an amazingly variable group of insects. Although they all possess a similar body plan comprising domed hemispherical bodies, short legs and bright colours, the colour and pattern variation within species is nothing short of astounding. In Britain the 2-spot ladybird, Adalia bipunctata, has both black and red ground colour (black is especially common in the North aiding heat absorption) together with spot patterns merging to form bands, blotches, rectangles and discrete blobs. In Japan and much of the Far East, the Asiatic lady beetle, Harmonia axyridis, has at least 25 variants with colour and patterning changing as the animal’s geographical distribution changes from East to West. Similarly, a look at any family photograph will reveal marked similarities and differences between close and extended family members. Physical factors such as hair colour, height and even relatively insignificant features such as whether or not the ear lobe is attached all contribute to a natural variation within the group. Even identical or monozygotic twins show differences in body structure through the effects of the environment or because of subtle changes during development. So how is this variation brought about? Perhaps the simplest explanation is that the two sets of traits (mothers and fathers) are somehow mixed and blended together like paints in a tin? Blending inheritance is a theory (never formally submitted scientifically) that the offspring of two individuals will be an ‘intermediate’ or a mixture of the two parents. A tall male and a small female, for instance, will give rise to progeny intermediate in height. In the absence of any alternative, the ‘blending’ theory was a commonly held belief in the eighteenth and nineteenth centuries and was accepted by Charles Darwin – although he was acutely aware of its shortcomings. Because

46

Genetic Variation within Populations

logically blending inheritance will ultimately remove variation – the offspring in future generations becoming more and more uniform, more similar, more ‘blended’? This is not what we see in nature - variation does not aggregate around the mean; variants are conserved. From his studies on domesticated animals Darwin knew that variation was heritable, but he was unable to suggest a suitable mechanism. The answer to this conundrum came six years after the publication of Origin of Species in a little-read paper published by Gregor Mendel, presented to his local natural history society in 1865. His particulate theory of inheritance postulated discrete characters or factors that are segregated during the formation of sex cells and recombined in the fertilised egg. We now recognise these factors as alleles. But it was only in 1900 when two biologists, the Dutch botanist Hugo De Vries and the German botanist Carl Correns, obtained hybridisation results similar to those of Gregor Mendel that Mendel’s work (and what we now refer to as Mendelian inheritance) became acknowledged within the scientific community. William Bateson (an English biologist working at Cambridge University) translated Mendel’s original paper into English (Bateson, 1902) and thereafter began to popularise his work. When observing variation, it is noticeable that the majority of physical characteristics show what is termed continuous variation, which is a smooth gradient of the character trait from a very large value to a small value along a continuous scale. The beak size of Galapagos finches demonstrates this as does other measurable characteristics of animals such as height, weight and colour intensity. This is the type of distribution most of us (and certainly the naturalists in the nineteenth century) are familiar with. Mendel however had the good fortune to look at plant characters that varied discontinuously, that is in separate categories. The pea flower was either purple or white, the seeds wrinkled or round, the pod inflated or constricted. Appearance of discontinuous or discrete forms (discontinuous or categorical variation) can also be seen in humans (evidence of albinism or in blood groups) where the character within the population is seen in two or more distinct forms. Discontinuous and continuous variation are examples of genetic variation that can best be described both in terms of phenotype and genotype. This type of variation within populations is further considered later in this chapter. The phenotype is the outward or physical appearance of the organism. Collectively, populations of animals and plants may be represented as frequency distributions (see Figure 3.1). Notably, a continuously varying character is often graphed as a bellshaped distribution with intermediate phenotypes more common than extreme ones. In contrast a discontinuously varying (or discrete) character appears as discrete blocks within a bar chart. The origins of the two distribution types are dependent upon the genotype (the genetic make-up) of the organism. For instance, in the Galapagos finch, G. fortis, we know that parents with larger beaks produce offspring with larger than average beaks (and vice versa). We also know that there is a precise relationship between the probabilities of a child having a particular blood group and the blood group of its parents.

Inheritance and Variation

Number of birds

15

50

AB64

30 20

5

10 0

0

15

20

Number of birds

AB68

40

10

AB99-02

AB03

15

10

10 5

0 30 Number of birds

47

25

5 0 15

AB04

20

BB04

10

15 10

5

5

Beak size (PC1)

2.25

1.50

0.75

2.25

1.50

0.75

0

-0.80

-1.50

0

-2.30

5

0

10

-0.80

15

EG05

-1.50

EG04

20

-2.30

0 35 30 25 20 15 10 5 0

Number of birds

0 25

Beak size (PC1)

Figure 3.1 Frequency histograms showing variation in beak size of G. fortis

(black bars represent mature males and arrows discontinuity between large and small beaks). Reproduced from Hendry et al. (2006) with permission.

If a character is determined by only a single allele, then the progeny within a population will conform to only a few discrete types. In Mendel’s pea flowers, for example, pure breeding Tall plants (the dominant character) crossed with pure breeding dwarf plants (the recessive) will produce a first (F1) generation of only Tall pea plants (the heterozygote, ‘Tt’). Thereafter crosses between these (F1) Tall plants will produce both Tall and dwarf offspring in the following (F2) generation in a ratio of 3:1 – see Figure 3.2. Similar results are seen with other traits in the pea plant such as purple and white flowers, round and wrinkled seeds.

48

Genetic Variation within Populations

Figure 3.2 Diagrammatic representation of Mendel’s early crosses using pea plants.

Inheritance and Variation

49

Figure 3.3 Continuous variation in human height (in metres). (A black and white version of this figure will appear in some formats. For the colour version, please refer to the plate section.)

Similar Mendelian genetics can be used to observe future F3 and F4 generations. Also, the monohybrid (single gene or single allele pair) cross illustrated in Figure 3.2 could be expanded to explore dihybrid (two genes) or trihybrid (three genes) crosses. But in characters such as human height, finch beak size or crop yield the observed variation could not be produced by a few alleles at selected chromosome loci. By increasing the number of alleles and the number of loci the frequency distribution becomes ever more continuous (Figure 3.3). Continuous variation is polygenic, that is controlled by many gene loci. Indeed, Mendel himself had noted (in his original 1865 paper) that multifactorial inheritance can generate a continuous frequency distribution, a point confirmed in genetics labs at the beginning of the twentieth century. Two biological phenomena have been alluded to above – variation (referring to differences within a species) and diversity (applied to differences between species). Origins of variation can be either genetic or environmental, but for any change to be heritable, a genetic influence is required. Short-term changes over the lifetime of an individual are termed ontogeny, whereas longer term changes, taking place over millions of years giving rise to distinct populations and even new species, are termed phylogeny. This distinction between the growth of the individual (ontogeny) and its ancestry (phylogeny) has been debated from the time of Aristotle onwards. Both these phenomena lead to changes within the individual genome, but over the longer time frame further genetic change may lead to an alteration of the entire gene pool. Genetic variation within a species is the ‘clay of evolution’, it is the substrate on which biological selection operates. As such it forms an important research focus for evolutionary biologists. This chapter deals with questions such as the origins of variation, the extent of genetic variation and patterns of inheritance (transmission of variation).

50

Genetic Variation within Populations

Early Ideas Regarding the Continuity of Life Strange hybrids of animals and human beings were a characteristic feature of ancient, and indeed more recent, mythologies. Reproduction was seen to be the dominant life force and by the nineteenth century over 300 theories of procreation had already been published. The Greek philosopher Aristotle was perhaps the first great scientist/ observer of the natural world. He produced many written works with around one quarter of those surviving relating to the Life Sciences. Three types of reproduction were known to Aristotle: sexual reproduction (in the ‘higher’ animals including some insects), regeneration (common in marine animals) and spontaneous generation (as seen in ‘vermin’, fleas, mosquitoes, maggots and the like). His main contribution to the subject of reproduction though was the concept of potentiality, the notion that an embryo is not complete and preformed but somehow acquires its potential during development. Male and female contributions are also specified. The female provides the basic matter or substance of the egg, the male provides the ‘form’ or the pattern in the seminal fluid – hence the idea that the role of the male in reproduction is somehow more significant than that of the female – an idea that persisted well into the nineteenth century (refuted only when Mendel demonstrated the equal contribution of the sexes). During the eighteenth century, contrary to the idea of potentiality, the theory of preformation was proposed. The theory of preformation supposed that an embryo arose complete in every detail (as an organism in miniature). The male was said to implant a preformed ‘homunculus’ within the body of the female. This theory was not dependent on what was seen but what was believed to exist. Following the results of further embryo studies however, a later theory of pangenesis supplanted the theory of preformation. The homunculus theory fell into disrepute and the doctrine of pangenesis was put forward to explain heritable changes. The assumption here was that somehow invisible particles or ‘gemmules’ were transferred (in the blood) from all the organs of the body to the sex cells. Thereafter the fertilised egg cell (zygote) redistributed the particles to the developing embryo and eventually to the adult form. The pangenesis theory has long-lasting consequences, for instance, the notion of a blood line (still common in animal breeding circles) and terms such as ‘blue blood’ or ‘blood relatives’, which continue in everyday speech.

Biological Inheritance and the Work of Gregor Mendel Charles Darwin knew nothing of modern genetics. Yet a knowledge of variation was central to, and a prerequisite of, his theory of evolution by natural selection. The first two chapters of his major work Origin of Species (1859) illustrate this. Chapter one is called ‘Variation under domestication’ and Chapter two ‘Variation under nature’. He published 19 books in all, drawing material from his personal observations of many kinds of animal and plant. He studied barnacles, living and fossil, orchids and

Biological Inheritance and the Work of Gregor Mendel

51

primulas. Darwin studied variation to explain diversity, but the (genetic) nature of variation was unknown to him. Between 1856 and 1863, Gregor Mendel, monk and latterly abbot of the monastery of St Thomas, Brunn, cultivated more than 33 000 plants and carried out dozens of experiments on the inheritance of different biological traits in the garden pea, Pisum sativum. He examined seven different physical traits or characteristics: v v v v v v v

seed shape (round or wrinkled) seed (cotyledon) colour (yellow or green) flower colour (purple or white) pod colour (green or yellow) pod shape (full or constricted) pod & flower positioning (axial or terminal pods and flowers) stem length (tall [6 feet] or dwarf [1 foot])

Mendel’s aim was to investigate inheritance of these characteristics and to look for mathematical regularities. Problems with blending (a theory of inheritance where offspring appeared intermediate in appearance to their parents) had been known for many years. If all offspring were some sort of parental blend, then after only a few generations much variation within the population would be lost and individuals would all tend towards an ‘average’ appearance. Common sense told Mendel, Darwin and others that this was not so. Individual animals and plants retained their distinctive characteristics despite evidence to the contrary from human beings (in skin tone, for example, many children are in fact a blend of their parents’ skin colour). The sensible approach therefore was to cross pure strains of his plants and observe the hybrids formed. No doubt Mendel must have considered breeding animals or other plant types, but pea plants suited him well. They were small, easy to cultivate and had many clearly defined, heritable differences; varieties of pea were collected and sold by gardeners and seedsmen. In addition, pollination was easily controlled, and the plant had a short growing time. Earlier studies had indicated to Mendel that the pea plant usually reproduced by self-fertilisation, but outcrossing could be achieved by simply removing the stamens (together with developing pollen grains) from the plant before the pollen was shed. The carpels (female parts of the flower) could then be pollinated by brushing with mature pollen grains taken from another plant. Upon analysis of these and many other results, Mendel derived several important conclusions: v He supported the view of particulate inheritance. The ‘factor’ determining inheritance was solid and relatively immutable, certainly not a result of blending. v Factors could be hidden but not destroyed. If both factors are present in the same plant, only one is expressed – the one that appears in the organism is termed dominant, the one that is masked is the recessive factor. v Factors occur in pairs and during gamete formation they are separated or segregated so that only one factor appears in the gamete.

52

Genetic Variation within Populations

v Factors segregate independently of one another. When considering several traits together, gametes produced may contain any random selection of (single) factors. v Factors recombine in the zygote at fertilisation restoring the pair. Mendel’s work was read to the local natural History Society in Brunn in 1865 and was published in 1866. There it remained undiscovered for 34 years. Mendel has, rightly, been called the founder of modern genetics. His methodical planning, mathematical analysis and quite outstanding insight has provided the platform on which all genetic analysis stands. Modern terminology has changed of course. Mendel’s factors are now recognised as alleles – alternative forms of a gene. Allele pairs may either be homozygous (alike) or heterozygous (differing), and, as they separate and recombine, they form the genotype or genetic constitution of the organism. Only certain alleles are expressed, and these contribute to the organism’s phenotype (its outward physical appearance). In practice though many alleles are linked together on the same chromosome (several thousands of genes of course but only tens of chromosomes). In theory therefore, all alleles on that chromosome will be transmitted together as a set. In practice, alleles are separated due to crossing over (as chromatids break and recombine). The nearer alleles are to one another (the closer their loci) the less likely they are to segregate. These events introduce an important exception to Mendel’s Second Law. Mendel also considered patterns of inheritance. He went from considering inheritance of a single characteristic (a monohybrid cross) to looking at the transmission of two independent traits (a dihybrid cross). For example, plants with wrinkled, yellow seeds could be crossed with those of round green seeds. As our ‘parents’ here are double homozygotes (rrYY and RRyy, respectively) then the resultant F1 hybrid will be the (double heterozygote) RrYy. Further genetic crosses of the F1 hybrids yield the familiar 9:3:3:1 ratio found in the F2 generation. To confirm independent segregation of alleles Mendel also undertook the arduous task of carrying out a trihybrid cross – the transmission of three independent traits. Theoretically a 27:9:9:9:3:3:3:1 ratio is expected (obtained by multiplying three independent 3:1 ratios). The excellent agreement of Mendel’s observed and expected results finally confirmed his views on particulate inheritance, segregation and assortment. Perhaps not surprisingly, this final activity for Mendel required ‘the most time and effort’. With many thousands of genes and relatively few chromosomes (numbering somewhere between 10 and 50 in most animals and plants) many genes must occupy sites on the same individual chromosome. Therefore, patterns of inheritance should differ from the standard Mendelian ratios. And this indeed is the case. If pure breeding purple/long-flowered pea plants are crossed with pure breeding red/round plants, the following (F1) generation should be all heterozygotes. However, in the early 1900s, William Bateson (inventor of the term ‘genetics’) and R. Punnett (of ‘Punnett square’ fame) found this not to be the case. In their F2 generation Bateson and Punnett noted the predominance of purple/long and red/round phenotypes – the original (F1) lines. They suggested perhaps some form of ‘coupling’ between these

Mapping the Genome

53

genes. Genes located on the same chromosome and transmitted together are said to show complete linkage. At the beginning of the twentieth century the chromosomal theory of inheritance was developed suggesting that Mendelian factors, the genes, were carried on chromosomes. Later work by Thomas Hunt Morgan on the fruit fly, Drosophila, confirmed this general conclusion – genes could indeed be coupled or linked during transport into the sex cells. Morgan’s team began mapping chromosomes showing how genes for characteristics such as white eye, miniature wings and yellow body are arranged linearly on the chromosome occupying limited but specific regions or loci. A view anticipating by several decades later work showing genes to be regions of chromosomal DNA.

Mapping the Genome Bateson and Punnett discovered that, on occasion, sweet pea plants failed to show independent assortment (during gamete formation the alleles segregate or separate independently of one another so that, theoretically, any member of the pair can appear in any gamete). Progeny did not follow the expected Mendelian ratio. Thomas Hunt Morgan noticed similar deviations when looking at inheritance patterns in Drosophila. The conclusion was that some form of physical ‘coupling’ existed between alleles; that different traits may be located on the same pair of homologous chromosomes. Morgan also suggested that when homologous chromosomes come together during meiosis, chromosomes may exchange genetic material thus providing (recombinant) cross-over products. Under the microscope chromatid interactions are seen (Figure 3.4) with chiasmata (the point at which paired homologous chromosomes remain in contact) visible. Further quantitative studies allowed a cross-over value to be determined: crossover Value ¼

Figure 3.4 Crossing over in Zea mays.

No:recombinants  100% Total no:offspring

54

Genetic Variation within Populations

Morgan discovered that the proportion of recombinant progeny varied considerably. He surmised that differences in cross-over values reflected actual distances separating genes on a chromosome (recombinants are more probable the greater the distance apart). Morgan’s student, Alfred Sturtevant, enhanced the model linking proportion of recombinants to gene loci. The per cent recombinant value was then used to determine linkage maps. It was decided that one genetic map unit (mu) was to be the distance between genes for which 1% of the products of meiosis were recombinant. 1mu 1mu

¼ RF ðrecombinant frequencyÞ of 1%ð0:01Þ ¼ 1 cM ðcentimorganÞ

A standard approach used in linkage analysis is the Three Point Test Cross (triple heterozygote to a triple recessive). Strange cross-over inconsistencies together with cytological evidence demonstrate the existence of double cross overs. And cross overs in one region may well affect the probability of cross over in an adjacent region, a phenomenon known as interference. Genetic interference can be quantified by calculating a coefficient of coincidence: Coefficient of coincidence Interference

¼

observable frequency of double recombinants expected frequency of double recombinants = 1  coefficient of coincidence

(if there are no double recombinants, then coefficient of coincidence = 0). In summary, crossing over: ✓ ✓ ✓ ✓ ✓ ✓ ✓

takes place in prophase 1 of meiosis where chromosomes shorten & fatten chromosomes associate in homologous pairs forming bivalents each chromosome is made up of two chromatids chromatids wrap around each other chromatids repel but are joined at certain points (called chiasmata) chromatids break and recombine with other (non-sister) chromatids

Genomics is that branch of biology that attempts to understand the molecular organisation of the entire genome (that is, the information content and the gene products that the genome encodes). (Genome = entire complement of genetic material in a chromosome set.) Genomics can be divided into two basic areas of study: 1. structural genomics characterising the physical nature of the genome and 2. functional genomics looking at the range of transcripts produced by a given organism (its transcriptome) and the set of encoded proteins (its proteome).

Mapping the Genome

55

Figure 3.5 Transmission of an X-linked genetic disorder (Duchenne muscular dystrophy). Duchenne muscular dystrophy (DMD) is characterised by a wasting and degeneration of muscle tissue due to the lack of the protein dystrophin. (Incidentally, dystrophin is coded by one of the longest human genes, 2.4-Mb in length). This is a sex-linked condition with the mutation appearing on the X chromosome. As a consequence, this condition is usually found in males (no corresponding allele to counteract its effects on the male’s Y chromosome).

The major aim of structural genomic analysis is the elucidation of the entire DNA sequence of a given species. Comparative genomics on the other hand compares the genomic sequences of different species with a view to understanding the course of evolution. It transpires that within species there is considerable synteny (conserved gene location within large blocks of the genome). In contrast to the detailed experimental work carried out by Mendel, the study of human genetics has been largely observational. Patterns of inheritance are generally derived by inference. The basic tool of the human geneticist had been that of pedigree analysis. Family pedigrees are presented, in diagrammatic form, in such a way as to make clear the transmission of characteristics (often genetic defects) between parents and offspring (see Figure 3.5). However modern advances in human gene sequencing are revolutionising human and medical sciences. To explore whole genome sequencing, tentative discussions took place in 1985 at the University of California, Santa Cruz, regarding the possibility of constructing a complete genome sequence for humans. Hitherto, DNA sequencing techniques had been developed in the mid-1970s by workers such as Ray Wu at Cornell, Fred Sanger at the MRC Centre, Cambridge and by Walter Gilbert and Alan Maxam at Harvard. Early work centred on specific regions of DNA, often specific genes, using restriction enzymes (discovered in the late 1960s) to ‘cut’ the DNA in a specific way. The selected DNA pieces were then incorporated into a vector (often a bacterial plasmid), which was introduced into a host cell (again bacterial) and cloned on a

56

Genetic Variation within Populations

suitable medium. In this way an appropriate quantity of the desired gene was obtained. A labelled probe (generally a single-stranded piece of complementary RNA or DNA) was then used to identify the DNA region under consideration. Then, using the polymerase chain reaction (PCR) process, DNA was heated to separate its two strands (the chromatids), and ‘free’ nucleotide bases (A, C, G, T) were used to re-form the original DNA. In this way, a physical mapping and ordering of DNA fragments had been completed. The move from manual sequencing methods (as employed initially by Sanger and his colleagues) to automated sequencing in the 1990s allowed molecular biologists the luxury of attempting to sequence an entire genome. In the early stages of whole genome analysis, a technique called shotgun sequencing allowed relatively small bacterial genomes to be sequenced. This approach (very briefly) involved: v Shearing DNA into random pieces v For each piece reads (approximately 500 bp) were identified from each end v A computer program assembles the multiple, overlapping ends or reads into a contiguous sequence (or contig) v Contigs are placed together to elucidate the entire genome This strategy was used to sequence the genome of the bacterium Haemophilus influenzae in 1995. The first eukaryotic genome (unicellular yeast cell, Saccharomyces cerevisiae) was sequenced the following year and the first multicellular organism (the nematode worm, Caenorhabditis elegans) in 1998. Of the other model organisms, the fruit fly genome (D. melanogaster) was sequenced in 2000. The Human Genome Project was launched in 1990 in the United States through funding from the National Institutes of Health and Department of Energy. Its goal was to map 90% of the millions of base pairs in human DNA and to identify all human genes from both a structural and functional point of view (the DNA of telomeres and centromeres was omitted from the project). The consortium was joined in 1993 by scientists in the United Kingdom at the newly opened Sanger Centre (funded by the MRC and Wellcome Trust). An international collaboration then followed with researchers using variations of the shotgun approach to identify the make-up of specific DNA sequences. An initial draft of the genome was published in the year 2000 and was finally completed in 2003. Interestingly competition from a privately funded company, Celera Genomics, producing a rival sequence, accelerated the end stages of the project. To summarise the main findings: the human genome was found to be smaller than expected with just over 20 000 genes uncovered; there were more repeated DNA sections than previously thought; and fewer than 7% of the protein families were vertebrate specific. Nearly half of the genes discovered had unknown function.

Origins and Maintenance of Variation Each organism has a unique genome derived from the parental segregation of alleles at meiosis together with a random combination of alleles at conception. A population

Mutation

57

therefore consists of a collection of unique genomes providing a gene pool with an inherent and measurable variability. Variation can be caused by internal or external events during the development of the organism. External, environmental factors can affect individual development. For instance, pine trees growing at the edge of a plantation produce needles that are smaller and thinner than those, more protected, trees at the centre, and human skin subjected to sunlight will begin to darken over time. Such externally induced effects are generally reversible and non-heritable. Individual variation can also be the result of internal, genetic factors. Genetic variation can arise through: 1. gamete formation at meiosis (for example, crossing over) 2. during recombination of homologous chromosomes at fertilisation 3. through mutation (alteration of the DNA strand) Also, phenotypic variation can occur during development as genes interact both with one another and their chemical environment. But to return to our three potential origins of genetic variation: 1. During metaphase 1 and metaphase 2 of meiosis (meiosis remember is a double division) chromosome pairs (bivalents) align themselves randomly across the equator of the cell attached to the spindle. The distribution of the chromosomes within each pair is random. The independent assortment and segregation of both chromatids (meiosis 1) and chromosomes (meiosis 2) will result in unique allele combinations within the gametes. Another feature resulting in genetic reassortment is the phenomenon of cross overs (see Figure 3.4). 2. Sexual reproduction also introduces an element of chance-mixing into chromosome combinations. Only one of two chromosomes from each parent will find its way into the gamete. If mating is random, then any one of many chromosome combinations (hence allele combinations) can occur in the zygote. This form of gene reshuffling forms a permanent impression on the genome of the zygote and is therefore heritable. No new forms are created, but this type of genetic variation gives rise to the continuous variation seen in features such as body mass, petal number, beak size of birds, etc. 3. Ultimately though, the origin of all evolutionary change is rooted in the creation of new forms and of new alleles – that is the production of mutations.

Mutation A mutation is any alteration in the DNA of an organism. This may encompass either the structure of DNA (that is the nucleotide sequence) or the amount of DNA. For a mutation to be heritable it must occur in the sex cells or gametes rather than the somatic or body tissue (for only mutations in the germ line are heritable). There are several kinds of mutation, each with their own characteristics. These can for simplicity’s sake be referred to as gene (or point) mutations and chromosome mutations, the former affecting a single nucleotide base, the latter changing the appearance of entire chromosomes.

58

Genetic Variation within Populations

In 1937 Dubine, a Russian biologist, discovered that up to 2% of fruit fly populations contained spontaneous mutations. This important observation was generally overlooked until about 20 years later as most biologists at the time viewed mutations as accidents and of no real significance. Work by Harris, Lewontin, Hubby, Stone and others in the 1960s established beyond doubt that mutation is a regular phenomenon, present in all species empowering variation and evolution. Point mutations come about through DNA replication and involve either gain (addition/substitution) or loss (deletion) of nucleotides. Substitutions are the commonest form of nucleotide change occurring at a rate of around one substitution every thousand million generations. In one example, a single point mutation in the β amino acid chain in human haemoglobin alters the haemoglobin molecule leading to the condition known as sickle-cell anaemia. A simple substitution of uracil for adenine in the nucleotide alters the triplet code from GAA to GUA. The amino acid valine is produced instead of glutamic acid. Therefore, an abnormal polypeptide chain forms leading to abnormally long haemoglobin fibres which, at times of low oxygen concentration, results in the ‘sickle’ shape of the owner’s red blood cells. The speed or rate of mutation varies dramatically according to the location, sex and type of base of the nucleotide. For example, mitochondrial DNA often mutates faster than nuclear DNA, while the DNA of the AIDS virus mutates faster than either of those. In mammals, mutations are six times more likely in sperm cells than in egg cells, and in 1977 the French biologist Lucotte suggested that mutation rate varied with the complexity of the organism. Thus, eukaryotes will have higher mutation rates than prokaryotes, vertebrates will have higher rates than invertebrates and ‘lower’ vertebrates’ higher rates than the ‘higher’ vertebrates and so on. Evidence for this point of view though is not very convincing. By and large, bacteria have lower rates of mutation than the Metazoa, and there is evidence that mammals have higher mutation rates than other animals. But the different ways of scoring mutation (based on morphological/physiological effects, mutation rate per gamete or mutation rate per genome) combined with factors that can elevate mutation rate (physical factors such as UV radiation, ionising radiation or chemicals together with transposable elements and ‘mutator genes’) all conspire to produce a confused picture. The DNA strand is not simply a string of triplets coding for specific amino acids. Up to 90% of its length can be made up of non-coding DNA. The internal structure of the gene is also remarkably complex consisting as it does of exons (nucleotide sequences forming amino acids), introns (sequences transcribed but not translated into amino acids) and flanking regions (important in regulating gene expression, see Figure 3.6). The probability of mutation appears to vary as to where genes are found in this complex structure. Regions of high guanine–cytosine content (called isochores) can be mapped onto the chromosome yielding specific gene assemblages. Any gene moving from one isochore to another will experience a shift in the rate and direction of mutation. A further class of highly variable and mutable DNA sequences are the variable number tandem repeats (VNTRs), a region used for genetic fingerprinting. VNTRs comprise highly repetitive regions of DNA within the genome. These sequences, sometimes called minisatellites, are small (usually 15- to 60-bp long),

Mutation

59

Figure 3.6 Detail of the protein-coding unit within a eukaryotic chromosome. (A black and white version of this figure will appear in some formats. For the colour version, please refer to the plate section.)

non-coding and suitable for use as genetic markers. VNTRs are however highly mutable with mutation rates of up to 15%. Naturally a common core is highly conserved, which will allow this material to be used as a genetic profile of the individual. The process of protein manufacture within cells can be conveniently divided into two stages:  TRANSCRIPTION: the copying or transcribing of the genetic code onto the (messenger) RNA molecule. This copy will pass out of the nucleus into the cytoplasm where it attaches to a ribosome. Studies of viral DNA infecting mammals revealed that viral DNA transcripts had little correspondence to the original DNA. The transcript was generally shorter in length with pieces removed. The RNA transcript therefore is processed in the nucleus before releasing different proteins. The transcriptional unit consists of those elements expressed in the formation of amino acids (exons) and intervening introns (not expressed). Flanking regions contain both ‘promoters’ needed for RNA polymerase recognition and ‘enhancers’ that control the amount and time of transcription.  TRANSLATION: the implementation or translation of the genetic code into a polypeptide molecule. Messenger RNA makes its way into the cytoplasm through the nuclear pores to ribosomes where chemicals, called initiation factors, help combine the ribosome subunits and assist in the attachment of mRNA. In 1927 Muller, an American geneticist, looked at the effect of induced mutation on fruit flies. Chromosome abnormalities can be caused by a wide array of physical and chemical disturbances as well as errors during meiosis such as non-disjunction (or separation) of chromosomes. Muller observed that rates of induced mutation were some 150x greater than control lines (with spontaneous, natural mutation). Chromosomes break naturally, but the frequency of this increases with ionising radiation, ultraviolet light and chemical agents such as mustard gas and acridine orange. The introduction of a virus into DNA may also have the effect of triggering mutation. Chromosomal mutations such as alteration in chromosome number or chromosome structure are responsible for a wide range of important human conditions. Common structural mutations include:

60

Genetic Variation within Populations

   

deletion (loss of a chromosome fragment during cell division), duplication (attachment of an extra segment on a sister chromatid), inversion (upside-down re-attachment of a chromosome fragment) and translocation (fragment joins a non-homologous chromosome).

Conditions caused by chromosome deletion include the mental retardation and small skull of ‘cri du chat’ victims, while translocation of fragments are implicated in several cancers including chronic myelogenous leukaemia. Translocation of a chromosome 21 fragment can produce Down syndrome. The evolutionary significance of structural mutations is to reduce the likelihood of cross overs during meiosis and increase the number of ‘normal’ gametes. Thereby the alteration (often an inversion) is conserved, the linked genes on the inversion (or altered component) being known as a supergene. Changes in chromosome number can occur spontaneously or can be induced to produce organisms for research or for the table! Many of our plant (and some of our animal) foods are variants resulting from this type of change. The number of chromosomes in a basic set is called the monoploid number (). Organisms with successive multiples of the monoploid number are called euploid. Euploid organisms that have more than two chromosome sets are called polyploid. Polyploid individuals are referred to (respectively) as triploid (3), tetraploid (4), pentaploid (5), hexaploid (6) and so on. We have already come across the haploid condition (n) seen in gametes. In most animals and plants, the haploid number and the monoploid number are identical (n = ). But in a few cases, for example modern wheat, haploid and monoploid numbers are different. Modern, agricultural wheat has 42 chromosomes. It is hexaploid so  = 42/6 = 7; its gamete however has 21 chromosomes and therefore n = 21. Polyploidy is common in the Plant Kingdom and an important feature in plant evolution. It is less common in animals, being seen in fish, amphibia and one mammal (a rodent, Typanoctomys barrerae). Within the set of polyploids we must distinguish between autopolyploids where chromosome sets come from the same species and allopolyploids where hybridisation allows the combination of chromosome sets from different (but closely related) species. In general, polyploid plants are often larger and have larger organs (e.g. larger fruits in strawberries) than their diploid relatives. Polyploid animals also occur but are much less common. The Salmonidae family of fishes (including salmon and trout) have twice as much DNA than related fish, while triploid oysters are more palatable than diploid forms. Human polyploidy is very rare and never results in viable offspring. Aneuploidy is the third example of (polyploid) chromosome mutation. The normal chromosome set is usually altered through abnormal separation of chromosomes (nondisjunction) during either mitosis or meiosis. As a result, the diploid (2n) condition either has additional chromosomes (e.g. Down syndrome, 2n + 1) or a reduction in chromosome number (e.g. Turner syndrome, 2n  1). These conditions are of especial importance to human health. Mutations can be induced by certain physical and chemical agents called mutagens. The high incidence of scrotal cancer in chimney sweeps was noted as

What Sorts of Genes Are Needed by Living Things?

61

early as 1775 – we now link the polyaromatic hydrocarbons in soot to this form of cell mutation. In the 1920s the mutagenic effects of X-Rays and Ultraviolet radiation was also demonstrated on fruit flies, providing discernible phenotypic (bodily malformation) and genotypic (chromosome) differences. But of greater evolutionary significance are the spontaneous mutations produced during DNA replication and repair. Mutation rates vary between organisms, between chromosomes and between loci, but with chromosomal mutation rates of 104 to 103 per gamete per generation and point mutations of around 105 to 106 per generation, the impact of this phenomenon on population change will be significant. Genes and chromosomes can mutate in either body (somatic) or reproductive (germinal) tissue and the biological significance of the mutation rests on whether the alteration affects only a small part of the body (a mutant sector) or whether the mutation affects the germ line and is passed on from generation to generation. It is in the study of population genetics where germinal mutations are of much greater significance. The action of mutation is to produce novel alleles, but to effect evolutionary change these alleles must persist within the population.

What Sorts of Genes Are Needed by Living Things? As we saw earlier in the chapter the science of genomics attempts to describe both the structure and function of genomes – that is the entire complement of genetic material within an organism. Moreover, DNA sequencing and genome analysis combined with the mathematical and investigative tools of bioinformatics have led to an explosion of genome sequencing projects including that of the human mitochondrion, reported in 1981, brewer’s yeast (Saccharomyces cerevisiae) in 1992, the first free-living organism (Haemophilus influenza) in 1995 and, most famously, the Human Genome Project finalised in April 2003. The Human Genome Project was an international and collaborative research exercise whose early results were published in Nature in 2001. The group recognised the existence of around 21 000 genes on 46 human chromosomes as an amazing blueprint for the establishment and development of every human being! The whole genome sequencing of the human genome has led inexorably to advances in health care in the emerging field of genetic counselling, predictive and preventative medicine. However, human genome sequencing also carries the attendant problems of social and ethical issues. As the technology advances and sequencing costs are reduced, then this area of study will increasingly be used in genetic scanning and manipulation of human and other genomes. As of 2011 there were nearly 3000 microbial (the majority were viruses) and 36 eukaryotic (half of which were fungi) whole genome sequences available. A complete genome sequencing has been carried out in several organisms including the nematode worm, Caenorhabditis, and the yeast, Saccharomyces. Results indicate that there is a division of labour between these genes with functional sets necessary for the correct functioning of the organism. The term proteome refers to the collection of protein-coding genes within an organism’s genome. In yeasts and

62

Genetic Variation within Populations

nematode worms genes can be classified into functional categories. These include (in order of size – largest first):       

cell metabolism nucleic acid metabolism protein folding and degradation transport and secretion signal transduction ribosomal proteins cytoskeleton

Living cells therefore seem to have a list of ‘core functions’ that need to be carried out. Over and above this, the remainder of the proteome results in species-specific and individual differences. A similar set of ‘core functions’ is seen in flowering plants, e.g. Arabidopsis thaliania, where it too has had its genome sequenced.           

22% of genes are involved in metabolism 15% in transcription 13% in defence 10% in secondary metabolism 8% in signal transduction 6% in growth 6% in energy liberation 5% in protein destination 3% in protein synthesis 2% in extracellular transport 2% in intracellular transport

Comparative genomics can provide a useful tool for the evolutionary biologist. Genome sequences have demonstrated both the unity of life (there is enough similarity to suggest that all life forms had a common origin) and the differences. For instance, it has shed light on what we mean by a species. Formerly hybridisation of DNA was the criterion for similarity (most bacterial DNA will hybridise provided the similarity in nucleotide sequence is >80%). But following the work of Carl Woese (1928–2012) bacterial species were defined by the variation of their (16S) ribosomal RNA (rRNA). Differences of around 2.5–3% are now considered to represent different species. And based on 16S rRNA analyses, Woese divided life on Earth not into five or six kingdoms as had been suggested previously, but into three domains. These were the Bacteria, Archaea and Eukarya. As species diverge, so do their DNA sequences. Pauling and Zuckerandl suggested that if such divergence occurred at a constant rate it would provide a molecular clock allowing scientists to date the splitting of distinct lineages and thereby providing a timeline for biological evolution. Some of the most interesting aspects of genomes are in the dynamics of gene expression patterns, the so-called functional genomics. But genomes also differ in

Genotypic and Phenotypic Variation

63

structure. Similarities and differences in genome sequences (structural genomics) can take the following forms: 1. 2. 3. 4.

differences differences differences differences

at the level of individual bases at the gene level in larger scale blocks in entire genomes

Duplication of genetic elements has been an important driver of evolution. Estimates vary, but the incidence of duplication is significant; in Arabidopsis thaliana for instance the genome contains more than 60% duplications. The evolution of human globins provide a good example of gene duplication. Genes for haemoglobin α and β chains are found on chromosomes 16 and 11, while other loci accommodate genes for myoglobin, neuroglobin and cytoglobin. Closely linked genes such as the α and β haemoglobins suggest a more recent divergence. Conversely less closely clustered genes diverged earlier. It appears neuroglobin split from the ancestral globin forms about a billion years ago with myoglobin having diverged around 500 million years ago and the α and β haemoglobin chains around 450 million years ago. Whole genome duplication has occurred several times in animals, plants and microbial forms. Expression of Hox genes determines bilateral symmetry and posterior/anterior axes with many metazoan animals. Hox genes reveal the duplications that have occurred during vertebrate evolutionary history. Amphioxus has one HoX cluster, humans possess four clusters and zebrafish seven Hox clusters. This has been interpreted as a series of large-scale duplications, 1 ➔ 2 ➔ 4 ➔ 8 (with zebrafish having lost a duplication, 8 ➔ 7).

Genotypic and Phenotypic Variation The phenotype of an organism describes its morphological, chemical or behavioural attributes and is produced by the action of genes in combination with the environment. Three qualities of a gene impact on the phenotype:  penetrance: the degree to which the condition controlled by the gene is manifest  pleiotropy: a single gene having multiple effects  epistasis: where a gene alters the expression of a second, independent gene Gene expression is the molecular process by which genetic information in the chromosomes is converted into protein molecules that determine the characteristics of the cell. Genes consist not simply of protein-coding units arranged linearly like beads on a chain, but are much more complex. In prokaryotes the nucleotide sequence matches the amino acid sequence in the polypeptide produced, an attribute known as colinearity. In eukaryotes the correspondence between DNA and the polypeptide chain is less precise. Gene expression is modified by the action of factors such as suppressors and modifiers. It is possible to measure gene expression by determining the penetrance or expressivity of that gene within the phenotype. These two terms,

64

Genetic Variation within Populations

penetrance and expressivity, are slightly different. Gene penetrance is defined as the percentage (or amount) of that particular characteristic present in the underlying genotype found in the phenotype: in other words what fraction of genes present in the cell are actually expressed. Expressivity refers to the degree to which genes are expressed (a pigment gene, for example, may not be expressed fully in the phenotype). When considering phenotype, it must also be remembered that genes interact. They interact with one another and they interact with their immediate environment. Gene interaction includes aspects of inheritance such as incomplete dominance, codominance, lethal genes, suppressors and duplicate genes (all of which can be studied in basic genetics textbooks). The interaction of genes with their physical–chemical environment has also been well studied using ‘temperature shock’ in fruit flies, light availability in plants and sex determination in crocodiles. Two generalisations are apparent: one is that a single gene can have multiple effects (pleiotropy) and the second is that a characteristic in the phenotype may derive from more than one gene. The latter phenomenon is perhaps easier to explain when considering, say, a biochemical pathway with several components each controlled by a single gene, but what about the converse? How are pleiotropic effects produced? In humans, sickle-cell anaemia has a multitude of symptoms caused by a simple base substitution. These include skin ulcers, heart failure, physical weakness and kidney failure. In Marfan syndrome mutation of a single gene causes near sightedness, malformation of the sternum and a weakened aorta. Similarly, dog and cat breeders have known for many years that breeding for one characteristic often brings with it a range of unwanted features, for example white fur/blue-eyed cats often have problems with associated deafness. Genes producing apparently unrelated effects are known as pleiotropic genes. There are two models of gene–environment interaction.  The first model postulates general genetic rules that are then qualified by more specific statements from the environment. So, for example, when constructing a model aeroplane, basic rules provide instructions to make a wing, fuselage, etc. Thereafter, specific statements regarding adornments and decoration determine the type of aircraft.  The second model is more deterministic in that the genes supply a detailed plan or blueprint, the environment serving only to modify slightly the overall pattern. In golden retriever dogs coat colour alleles are affected by nearby colour deposition alleles. The dominant form of the deposition allele promotes colour formation in the fur, the recessive form masks colour production. Therefore, genotypically a dog may be black or brown in colour, but if the double recessive form of the deposition allele occurs, then no colour is deposited, and the familiar golden retriever is the result. In the situation described, the expression of one gene is affected by a different gene at a different locus. This genetic condition is known as epistasis and accounts for much of the genetic interaction during development. For many years, biologists have noted the loose coupling of genotype and phenotype. As we have described earlier in this chapter, the correspondence or colinearity between genes and gene expression is very variable particularly in

Genes in Populations

65

eukaryotes. Organisms may look similar (phenotypically) but differ significantly in genotype; the convergent evolution (similar structures though different evolutionary trajectories) of both animals and plants provides many examples of this. Alternatively, organisms may differ in phenotype but have very similar genotypes. Examples here include the many polymorphic forms of animal and plant together with those species, like Amphibia and metamorphic insects, that change their body form during their lifetime. The two main reasons for this loose coupling of genotype and phenotype are:  that much of the DNA is never transcribed so that genetic and environmental changes will have little effect on phenotype;  regulatory genes (e.g. Hox genes) can dramatically alter developmental pathways and therefore eventual phenotype with relatively little genotypic modification. It has been stated before that the genome is the ‘clay of evolution’, the basis upon which a natural variation arises both in individuals and populations. Selection of course acts on individuals. But it is within the population that the macro changes occur and ultimately new forms arise.

Genes in Populations The marriage of Darwinism and Mendelism in the mid-twentieth century gave rise to the modern synthesis. Within this framework, individual biological disciplines such as palaeontology, taxonomy, biogeography, evolutionary theory and population genetics combine to emphasise the role of natural selection in evolution. A central tenet of the modern synthesis was that selection acts on the individual – but it is the population that evolves, not the individuals themselves. The essence of any evolutionary change therefore is an alteration in population gene frequency. The population is a basic unit of biology. It is not a static feature but a constantly changing entity shaped by its surroundings, both biotic and abiotic. A population of organisms can change both quantitatively (for example, numbers of individuals or population density) and qualitatively (as the characteristics of its constituent members change). A population is a group of organisms of the same species occupying the same habitat at the same time.

Populations can differ; for instance, a northern population of butterflies, such as the northern brown argus butterfly, Aricia artaxerxes, in the United Kingdom, contrasts with its southern counterpart by the presence or absence of a white spot on the upper wing. Similarly, plants such as the early purple orchid and rock rose show distinct northern and southern ‘races’ with zones of hybridisation in between. We also know that populations in the past can differ from populations we see today. Population biology looks at both geographical and temporal variation. Demography is a study of processes that determine population change. Individuals are born and they die; groups migrate into and out of populations. Thus, the growth potential of populations under

66

Genetic Variation within Populations

different environmental conditions is one measure of their evolutionary success. Although the term ‘demography’ is often restricted to human populations and human statistical analysis, the topic can also explore the statistics of sex and age to assess the mortality and natality rates of any biological population (this has been termed biodemography). Both Charles Darwin and his younger half-cousin Francis Galton noted that within animal and plant populations individuals ‘overreproduced’, that is, produced more individuals than is necessary to replace themselves. Populations increase in several ways, but population growth is generally geometric rather than arithmetic. In other words, growth accelerates due to the participation of all individuals (geometric) rather than the simple addition of numbers (arithmetic). But no population of organisms can continue to grow indefinitely. A variety of environmental factors (collectively labelled environmental resistance) will determine the eventual size of the group. Some of these external factors are density independent such as fire and flood; the effect on the genotypes of individuals is almost random. But others such as predation, food availability, etc. are density dependent and, as such, will selectively alter the eventual genetic make-up (the gene pool) of a population. This of course is a cornerstone of natural selection. Of course, a key element in population growth is reproductive capacity. A useful concept in this respect is that of a panmictic population, where the probability of mating is almost equally probable between any two individuals. Contrast this with a more structured population, say human social or geographical subgroups where mating is restricted by distance or social convention. A panmictic unit is one in which individuals are sufficiently close to allow the possibility of random mating with any other individual within that unit.

Variation within Populations The goal of population genetics is to understand the genetic composition of a population (its gene pool) and to describe the forces that impact on both individuals and the group as a whole. Until the middle of the twentieth century it was widely believed that genetic variation was limited in scope. Despite an impressive knowledge of human variability, the consensus was that animal and plant species were, by and large, genetically uniform. It was the work of population ecologists such as E. B. Ford (1901–1988) and his coworkers that brought the complexities of genetic variation to our attention. In the British Isles there has been a long tradition of natural history, resulting in the work of eminent naturalists such as John Ray, Gilbert White and Charles Darwin being placed in national and regional collections. And although animal and plant species have a generally uniform appearance, the existence of races, varieties, subspecies or ‘sports’ has been well documented. Intraspecific diversity has been recorded in a wide variety of animals and plants, for example in the banded snails Cepaea nemoralis and C. hortensis (see Figure 3.7). Several genes control the shell

Variation within Populations

67

Figure 3.7 Polymorphism in the banded snail. A composite image of polymorphism in the whitelipped banded land snail. Photo credit, Ian Alexander (A black and white version of this figure will appear in some formats. For the colour version, please refer to the plate section.)

appearance – genes for the ground colour, presence or absence of bands, spreading of the band pigment and colour of the aperture lips (black lips as in C. nemoralis and white lips as in C. hortensis). The ‘colour’ genes are multi-allelic and very often on the same chromosome. All in all, these snails have a complex genetic make-up. The (phenotypic) variation in colouring and banding patterns is generally referred to as a polymorphism with the appearance of morphs dependent upon the immediate environment, some colours and patterns providing camouflage others influencing albedo values (reflectance of light). But an interesting aspect of recent research (Silvertown et al., 2011) is that ‘the relation of colour and banding patterns of banded snails has indeed changed and that this has taken place at an astonishing speed’. Further details regarding banding in Cepaea are found in Chapter 4. Regarding global geographical variation of animal populations, we can safely say that the morphology, physiology and behaviour of these organisms appears well suited to those environments. A series of ‘ecological rules’ have been suggested to describe the distribution of different types or morphs:  Bergman’s Rule – related forms are larger in colder regions and smaller in warmer regions  Allen’s Rule – body extensions such as ears, tails and beaks tend to be smaller in colder regions

68

Genetic Variation within Populations

 Glober’s Rule – dark body pigmentation is more highly developed in warm, humid regions These generalisations (they are not strictly biological rules) can be useful when considering evolved, adaptive traits. A difficulty when discussing ecological variation lies in determining how much of the variation is genetic and how much is environmental (or perhaps the result of an interaction between the environment and development of the organism). Sorting out environmental effects (such as food, light, moisture, etc.) on the organism and weighing up the contribution of its genes requires sophisticated techniques. Fortunately, there are mathematical techniques (sometimes simply referred to as quantitative genetics) that allow us to test for heritability - that is how much of the phenotypic variation is caused by genetic differences. The assumption is generally that the phenotype (P) is the result of both genetic effects (G) and environmental effects (E). This assumption, of course, ignores the complex interplay between genes and environment. If phenotypic variation (Var P) = genetic variation (Var G) + environmental variation (Var E), then it should be possible to carry out controlled experiments to estimate the part played by each component. For example, genetic variation can be reduced to almost zero by using genetically identical organisms (e.g. cuttings taken from the same plant or the use of genetically inbred animals). When different environmental conditions are applied, the heritability (H2) of the condition can be measured: H2

¼

Var ðGÞ Var ðPÞ

So, when measuring the variation within a population (the total phenotypic variance) we consider two factors, the variance between the genotypes and the environmental variance. Genetic variance is determined: 1. through crossing experiments (producing homozygous lines from which heterozygotes are formed/then measuring the phenotypic variance within each heterozygote genotype); and 2. by considering genetic similarities between relatives (their genetic correlation). The remainder constitutes the environmental variance. If heritability represents the proportion of phenotypic variance attributable to genetic causes, then the ‘environmentability’ of a trait is its analogue – the amount attributable to environmental variation. Both heritability and ‘environmentability’ show important attributes: v They are population concepts, not referring to any individual. v They depend upon the environments that animals find themselves in (if the environments are much the same, then phenotypic differences will be mainly due to genes – heritability will be high).

Variation within Populations

69

v The concepts are abstract – they refer only to proportions and tell us nothing about the genes themselves. Heritability often arises of course when discussing the nature/nurture question. It can be described as: the amount heredity contributes to an individual trait. However, this is not quite accurate, a better definition might be: ‘heritability is the proportion of the total variation between individuals in each population due to genetic variation’; this number can range from 0 (no genetic contribution) to 1 (all differences on a trait reflect genetic variation). Following on from this definition: v The number does not apply to individuals – only to variations within a population. v This value is not fixed; differences among groups (in the range of genetic variation and environmental variation) will result in different estimates of heritability. From more than one thousand heritability estimates the average heritabilities were calculated. Heritability of physiological parameters such as oxygen consumption and resistance to heat stress averaged around 0.33 (33%), while structural features such as wing length and body size averaged around 0.46 (46%). These values are relatively high when compared with the heritability of traits connected with reproductive success (e.g. H2 values for fecundity, viability and survivorship average around 26%). These characteristics are so important that any genetic variability has been eliminated by natural selection. Advantageous alleles have been consolidated by repeated selection thereby causing the heritability to decrease. Table 3.1 shows percentage estimates of heritability for typical farm animals. Once again heritability is seen to be decreased in those traits linked to fitness and reproductive success. Table 3.1 Percentage estimates of heritability in UK livestock (various sources) Livestock

Character observed

% Heritability (H2)

Friesian cattle

White spotting

95

Butterfat content

60

Milk yield

30

Conception rate Pigs

White leghorn poultry

1

Back fat

55

Body length

50

Weight at 180 days

30

Litter size

15

Egg weight

60

Egg production

30

Body weight

20

Viability

10

Estimates of heritability apply only to the population under study in that environment.

70

Genetic Variation within Populations

The presence of variants in a population is termed polymorphism (many ‘morphs’ or forms) – that is two or more forms that are genetically distinct but still form part of the same species group. Within the phenotype, variation is observed at several levels:      

morphological differences physiological differences chromosomal differences biochemical differences molecular/immunological differences behavioural differences

And within the genome, variation can be described either as a DNA sequence polymorphism (that is the inherent variation in base sequences) or as a variation in allele frequencies (an allele being an alternative form of a gene). Population genetics deals with genotypic variation but, by definition, only phenotypic variation can be observed.

Variation between Populations Having already mentioned the banded snail earlier, there are two further classic stories of polymorphism within animal and plant populations. In the case of the peppered moth, Biston betularia, the black form or morph predominates in the industrial Merseyside area whereas the peppered form predominates in the more rural areas of North Wales. This occurrence is associated with a phenomenon known as industrial melanism and is controlled by very few genes. Several butterflies and moths are known to produce black or melanic forms. Bernard Kettlewell in the mid-1950s suggested that the proportions of dark and peppered forms observed was due to selective predation by birds. To test this hypothesis he pinned different moth variants to tree trunks and observed the responses of birds. Sure enough, the dark forms survived better on the soot-blackened (industrial) tree trunks and the peppered forms showed increased survivorship on the speckled, lichencovered trunks in more rural areas. One criticism of this work was that it was an artificial situation; moths normally rest (higher) elsewhere in the trees. But mark– release–recapture investigations provided equivalent results. Cyanogenesis (production of cyanide) is a phenomenon found in around 50 orders of flowering plants. Like most secondary plant compounds (phenols and mustard oils for example) these chemicals are defensive in nature. In the flowering plant bird’s foot trefoil (Lotus corniculatus), cyanogenic specimens are found mainly in woods and fields inland whereas the acyanogenic form is found nearer the coast. Cyanogenesis is controlled by two genes, one that controls production of a harmless sugar-cyanide compound and the other the production of an enzyme to release hydrogen cyanide

Variation between Populations

71

once the leaf is damaged. Natural populations are polymorphic for both genes but, once again, selective predation (this time by slugs) secures the survival of the cyanogenic forms in inland areas and acyanogenic forms in coastal zones. Slugs eating the (cyanogenic) trefoil leaf damage the cells, release the enzyme and are deterred from eating further. Now presumably the production of cyanide in the cyanogenic forms has an energy cost, so why is this condition predominant in inland areas but not near the coast? The answer lies in the distribution of the slugs. As any gardener will tell you, slugs do not like salt. And so, plants near the sea will not be predated by slugs, while those inland will almost certainly be attacked. Therefore, within a polymorphic population there will be differential survival (and therefore differential reproductive fitness) between those populations on the coast and those inland. In both these cases we are not looking at a conscious decision of individuals – it is a gradual shift in the character of the population. As the population begins to change (for example during the rise of the melanic form of B. betularia in the late nineteenth century), we are looking at a transitional polymorphism. Once the populations have become established, such as those in L. corniculatus, then we describe this stable genotypic frequency as a balanced polymorphism. Julian Huxley (1887–1975), grandson of ‘Darwin’s bulldog’ T. H. Huxley, coined the term cline to describe a continuous change in a phenotypic trait along a geographical axis. Clines may exhibit a gradual or graded change such as the increase in body size of many North American mammals and birds with increasing latitude; alternatively, change may be sudden, abrupt and discrete such as genetically distinct populations of house mice in Madeira or the different sized populations of dog whelks on the Welsh coast. Clines reflect an environmental gradient and can be explained using an adaptationist model. The larger size of North American mammals and birds could be ascribed to a reduced surface area/volume ratio and increased heat conservation. Dog whelk populations in Wales seem to be responding both to the physical forces of wave action (smaller animals on exposed shores) and to predation pressure (thicker shells on sheltered shores with more predators). John Endler (1947–present) described four ways in which a cline might form: 1. random genetic drift creating a difference in allele frequencies within isolated populations; 2. continuous environmental gradients such as temperature, altitude or humidity gradients; 3. spatially discontinuous changes in environment (e.g. island and mountain top populations) and 4. establishment of contact between genetically distinct populations. To this list might also be added changing predation pressures and human intervention. Clines demonstrate that strong localised selection pressures can produce local adaptations irrespective of gene flow, but how does genetic change in populations arise?

72

Genetic Variation within Populations

Population Genetics In 1908 Messrs Hardy and Weinberg independently described mathematical rules by which genotype frequencies behave in the absence of selection. The genetic puzzle up until this point was to describe how genetic variation was maintained in a population. Many alleles are deleterious and should, perhaps, be removed from the population – yet we often see a remarkable genetic constancy. At the turn of the twentieth century, it was argued that dominant alleles should be more common in the population – in the ratio three to one in fact. The existence of dominant alleles at relatively low frequencies suggested that Mendelian dominants and recessives were not segregating properly. But Godfrey Hardy in England and Wilhelm Weinberg in Germany (a third, independent discovery was also made by the Russian geneticist, Chetverikov) disproved this supposition by demonstrating that allele frequencies are not dependent upon dominance or ‘recessiveness’ but remain essentially unchanged from generation to generation. The Hardy–Weinberg principle states that: ‘in a sexually reproducing population the frequency of both dominant and recessive alleles will remain constant provided that certain conditions are met.’

The conservation of gene (and genotype) frequencies is one of the most important concepts in population biology and responsible for our further understanding of human diseases (carrier individuals) and selective breeding programmes (heterozygote advantage). The Hardy–Weinberg principle can be restated as a formula. If we imagine two alleles (A) and (a) at a single locus and a population of individuals mating at random, then the expected allele frequencies and genotype frequencies can be calculated using simple algebra. If we say the frequency of the dominant allele (A) = p and the frequency of the recessive allele (a) = q, then the following two formulae represent the Hardy– Weinberg equations: p þ q ¼ 1:0 ðto describe allele frequenciesÞ p2 þ 2pq þ q2 ¼ 1:0 ðdescribing genotype frequenciesÞ The gene pool represents the sum of all genes in the reproductive cells of a population. It can be regarded as a genetic reservoir from which samples are taken at random to create the next generation. We have seen that given certain conditions, the frequency of alleles and the frequency of genotypes in the gene pool remain constant. What are these ‘certain conditions’ and how might they cause deviation from the Hardy–Weinberg equilibrium? Certain assumptions must be made if we wish to hold to the Hardy–Weinberg equilibrium: v mating has to be random and

Population Genetics

73

v there must be no  selection of alleles either positive or negative  migration of alleles into or out of the population  mutation creating new alleles  genetic subdivision of the population  linkage disequilibrium disrupting the predicted occurrences of linked genes Deviations from the Hardy–Weinberg equilibrium often occur when there is nonrandom mating. For instance, it can transpire that individuals mate with others closely related to themselves (inbreeding). As a rule, inbreeding reduces heterozygosity in the population increasing the proportion of homozygous individuals. And, as many of the deleterious genes are recessive, an increase in homozygosity will increase the frequency of the homozygous recessives thereby increasing the probability of disadvantageous conditions in the population. Population variation is reduced; systematic inbreeding between close relatives will eventually lead to complete homozygosity. A second form of non-random mating occurs when individuals choose a mate not based on genetic relationship but because of their resemblance and mating preferences (assortative mating). Bias towards choosing a mate like one’s self is termed positive assortative mating; mating with unlike partners is called negative assortative mating. Human mate choice is often put down to positive assortative mating!

4

Natural Selection and Adaptive Change

In chapter 3 of Origin of Species, Charles Darwin muses on the reproductive rate of elephants: The elephant is reckoned to be the slowest breeder of all known animals, and I have taken some pains to estimate its probable minimum rate of natural increase: it will be under the mark to assume that it breeds when thirty years old, and goes on breeding till 90 years old (something of an overestimate of its longevity), bringing forth three pairs of young in this interval; if this be so, at the end of the fifth century there would be alive fifteen million elephants descended from the first pair.

Darwin was of course making the point that no animal species, or that of any other organism, ever achieves its full reproductive potential. This is vividly the case with organisms that produce significantly more potential offspring per generation; for instance, cod who lay more than a million eggs over their lifetime or the vast clouds of fungal spores. There is an enormous overproduction of offspring and yet the population numbers of most established species are roughly stable from one generation to the next. One frequent limiting factor on population growth is that of natural resources. Both Darwin and Alfred Russel Wallace acknowledged the influence of the Rev. Thomas Malthus in inspiring their respective theories of natural selection as the principal evolutionary mechanism. Malthus’s An Essay on the Principle of Population was first published in 1789 and considerably revised in 1803. It was this second edition that both Darwin and Wallace read. Briefly and brutally Malthus’ principle was that any attempt to ameliorate the lot of the (human) lower orders was doomed to failure because they would reproduce to the limit, beyond the available food supply – the poor would always be with us. The only hope was to encourage restraint in reproduction. Malthus’s insight was, of course, not necessarily true of humanity, but both Darwin and Wallace saw it as a spur in the development of their theory. In summary, all organisms are capable of exponential population growth, but this is never achieved. Typically, populations of organisms in the wild are stable or fluctuate in numbers well below their reproductive potential. Some go extinct, others may expand rapidly into a new environment, but stability is the norm. Only a small proportion of potential or actual organisms survive to reproduce. Both Darwin and Wallace acknowledge their debt to Malthus (and may have also been inspired along similar lines by Adam Smith’s (1776) An Inquiry into the Nature

Natural Selection and Adaptive Change

75

and Causes of the Wealth of Nations). But Darwin also records the inspiration of Archdeacon William Paley’s (1802) Natural Theology; or Evidences of the Existence and Attributes of the Deity Collected from the Appearances of Nature, which was probably read by Darwin while he was still an undergraduate. Paley’s ideas set up a challenge to anyone attempting a natural explanation of evolutionary change. His thesis is that organisms (particularly humans and other animals) are machine-like, and that their anatomy and physiology are evidences of deliberate design without evolution and thus proof of the existence of a designing God. Darwin saw it as his greatest challenge to explain the origin of adaptation by natural processes. We have already seen the first stage of his argument leading to the inference that only a very small proportion of organisms in any generation survive to reproduce. The second stage concerns the nature of natural populations. There is variation between individuals in almost all populations; the exception would be a population of identical clones reared under identical conditions! Some of that withinpopulation variation is heritable; a fact known to Darwin and Wallace from agricultural breeding as well as other examples seen in nature. This was further amply documented by genetic studies later in the twentieth century. Of that heritable variation it can also be demonstrated that at least some of it represents a difference in ‘fitness’, meaning the potential to survive and reproduce in a particular environment. Charles Darwin demonstrated that in any population of organisms constrained by natural resources the fittest individuals will survive differentially and thus pass on the heritable traits that confer increased fitness to the next generation. Over time those traits will increase in frequency in the population (given a constant environment). Any new and fitter feature must therefore appear in the population before it can be tested by natural selection. Wallace’s view was that heritable variation in populations was ‘random’ with respect to the direction of natural selection, which he saw as the cause of all adaptive change with one startling exception - the evolution of the human mind could not be explained by natural processes. Possible rivals to randomness include two factors proposed by Jean de Monet, Chevalier de Lamarck (1744–1829):  The first of these, thought by many to be the entirety of Lamarck’s theory, is the ‘inheritance of acquired characters’. This hypothesis attempted to explain the origin of adaptive features in an organism and its descendants. An animal, it suggests, develops a new characteristic because of its own striving in response to the environment. The lengthening of the neck in the giraffe to reach higher in browsing trees is the school textbook example. Thereafter, the acquired character becomes a hereditary one, so that the giraffe’s offspring, when mature, will have the long neck of their parent, even if there is no further effort on their part. This principle was then extended by Lamarck to include the origins of new organs arising from the animals’ environmental needs, and the ‘use and disuse of parts’ resulting in their enlargement or reduction, respectively.  But all these facets of ‘the inheritance of acquired characters’, which even Darwin was open-minded about, are not the core of Lamarck’s theory. Lamarck’s core idea,

76

Natural Selection and Adaptive Change

later termed ‘orthogenesis’ was the notion that in all organisms there was an innate drive to evolve from one generation to the next. Lamarck believed, as did many other naturalists into the nineteenth century and beyond, that all organisms (especially animals) could be arranged in a linear series, the scala naturae (ladder of nature) from most primitive at the bottom to the most advanced (inevitably Homo sapiens) at the top. Orthogenesis is the theory that creatures are programmed to ascend this ‘ladder’ and evolve along some predetermined track over the generations. It was very popular, particularly among palaeontologists, at the end of the nineteenth century (although Lamarck himself conceded later that there must be at least two animal scalae). Notice that both ‘acquired characters’ and orthogenesis are theories that postulate an innate component in the determination of evolutionary direction. In the first, the animal itself can be thought of as contributing a directional component to the evolution of its descendants; in the second, direction is entirely innate. In the early years of the twentieth century another type of innate evolutionary force was suggested by pioneer geneticists, notably Hugo de Vries, H. L Johannsen and William Bateson. Evolutionary change, including the origin of new species, occurred by ‘saltation’, a sudden radical ‘mutation’ (from the Latin saltare to leap about) between one generation and the next. Natural selection, if it acted at all, merely tidied up lethal mutants. It was the singular triumph of the ‘Modern or Synthetic Theory’ with its development of population genetics that reconciled Darwinism and Mendelian genetics.

Natural and Artificial Selection Organisms reproduce prolifically; often more than the capacity of the environment to sustain them. Under these conditions, of increasing population size, there will naturally appear a competition for resources such as potential mates, food and shelter. Which organisms survive and reproduce and which ones die of starvation or fail to reproduce is not random; a range of factors will contribute to their reproductive success. The process by which these factors govern survivorship is known as selection. An environment is said to select for those factors (and those individuals) favourable to success and to select against other factors that are less favourable. The language unfortunately implies a choice – of which there is none. Selection in this sense is a consequence of an interaction between living things and their environment. The survivors of this process are now thought of as being more closely adapted to their environment, that is, they can survive more readily. For selection to occur four conditions are necessary: v The units describing the trait (to be selected) must be variable. v This trait must vary within the parent population. v There needs to be an association between the trait and the organism’s reproductive success. v The trait must be heritable for selection to be effective.

Natural and Artificial Selection

77

Although ‘survivors’ of the selection process are better adapted, simple natural selection is a dynamic, active process. Genes that produce better adaptations become more frequent over time. A simple example will explain. The beach deer mouse, Peromyscus polionotus, lives in the Southeast of the United States in grass-covered sand dunes or scrub areas nearby. They have variable colouration: those living in darker soil burrows are coloured predominantly brown, and those in lighter, sandy areas are predominantly white. A good example of adaptation to their environment perhaps; but what of the mechanism, how did this situation arise? In this instance one can quite readily envisage a natural selection by predation. Lighter mice on darker soils are more apparent to a predator than the same mice on lighter sandy soils. This hypothesis was put to the test by Donald Kaufman at Kansas State University who introduced a hungry predator (an owl) into outdoor cages with equal numbers of light and dark mice. Perhaps unsurprisingly, the white mice on light soil survived better than the dark ones (and vice versa). The superficial explanation for the differential survival is, of course, improved camouflage providing a selective advantage. However, to gain a deeper understanding of the process we need to examine the genetic composition of the beech deer mice population. For several variants of the coat colour allele exist. And the shift in allele frequency (from light to dark or from dark to light depending upon environment) is dependent upon specific allele forms or allelomorphs being removed from the population – in this case by a hungry owl! To generate an adaptation to the environment the following are required: v v v v

a genetic variability within the population; the ability to pass on the specific variation to the next generation (heritability); a developmental link between genotype and phenotype; the capability to ‘translate’ adaptive genes into new offspring.

By contrast, artificial selection is the deliberate selection of organisms by humans for their benefit. The term artificial however is something of a misnomer. The selection is real, but the agency is not the unthinking natural environment but the purposive hand of human beings. The term, of course, is most closely associated with animal and plant breeding – the systematic reproduction of favoured types. The hope is that through selective breeding the next generation will contain more of the desirable characteristic. Thereafter if selected individuals are also mated then a greater and greater frequency of that characteristic (be it milk yield, thick fur, quantity of seeds, palatability of fruit) will be found in subsequent generations. Selective breeding is evidenced through an exploration of stock records at Smithfield market London (increase in carcass weight of cattle or the fleece weight of sheep over the past 200 years) or global crop yields for rice and wheat. Similarly, cat and dog breeders, pigeon fanciers (breeders) and rose growers all tell of deliberate mating of organisms with desirable characteristics causing replication of those characteristics in the offspring; indeed, some new ones! Artificial selection is also seen in the fruit and vegetables in our kitchen. A range of edible brassicas have been produced from the

78

Natural Selection and Adaptive Change

Cabbage - supression of intermode length Broccoli - suppression of flower development

Cauliflower - sterility of flowers

Kale - enlargement of leaves

Wild mustard

Kohlrabi - enhancement of lateral meristems Figure 4.1 Popular cultivated vegetables derived from wild mustard. Credit Creativ Studio Heinemann / Getty Images (cauliflower); Avadhesh Maurya / iStock / Getty Images Plus (mustard flower); Creativ Studio Heinemann / Getty Images (kohlrabi); Joff Lee / The Image Bank / Getty Images (kale); Axel Göhns / EyeEm / Getty Images (cabbage); Patricia Soon Mei Yung / EyeEm / Getty Images (broccoli).

rather uninspiring wild mustard (Figure 4.1), an aggressive weed only edible in small amounts in the young stages. Fish farming, or aquaculture, is a centuries old source of food production with European and Asian fishponds providing a plentiful and reliable source of nourishment and raw materials. But with acute land shortages the only way to increase production from fish farms is to increase yield. Fish yield can be increased either by altering the environment (antibiotics, food, temperature and water quality) or by manipulating the genome of the fish – that is growing genetically improved animals. If both approaches are employed, yields can improve dramatically. There are several breeding programmes that can be used to improve fish stocks genetically. Traditional approaches include selective breeding and cross-breeding, while more recent approaches include the production of sex-reversed brood stock and chromosomal manipulation.

Natural and Artificial Selection

79

 Traditional selective breeding programmes select only the most promising fish (size, colour, egg mass, etc.) from a fish population from which to breed. The resulting offspring will then embody many of these favourable traits.  Cross-breeding however attempts to find mating combinations from different fish populations; such offspring should then exhibit hybrid vigour. Cross-breeding programmes are a tried and tested technique, but results are unpredictable unless the combinations have been evaluated beforehand.  Sex-reversed programmes use the fact that one sex may be more valuable than another. And so, the sex of a population or brood may be reversed through the use of male and female sex hormones (oestrogens and androgens). For instance, female sturgeon are more valuable than males because they produce the caviar (eggs), while male tilapia are more useful because they grow twice as fast. One of the major goals in tilapia farming is to prevent reproduction, and this is best accomplished through generating a monosex male population.  Chromosomal manipulation is a recent technique that produces sterile fish. This is done to induce sterility in species that mature before they reach market size or to comply with legal restrictions that apply to growing exotic species (whose culture otherwise might be illegal). For example, grass carp culture is only legal in much of the USA if the aquaculturist raised sterile individuals (generally triploids). The most common form of chromosomal manipulation (producing triploid individuals) is to treat the newly fertilised eggs with heat or pressure in order to ‘shock’ them into preventing the second polar body nucleus from leaving the egg. In this way the zygote will contain a haploid sperm nucleus, a haploid egg nucleus and a haploid second polar body. Thus, producing a triploid, and therefore sterile, zygote. In order to characterise the theory of natural selection, we are now able to provide a simple description of this process: In the competitive struggle for existence better adapted individuals increase in frequency in a population with time, this differential increase we call natural selection.

Many authors cite something like this, backed up by statements regarding the overproduction of individuals within a population and differences in heritable adaptedness among individuals. Put simply, we can re-state the principles of natural selection as follows: 1. There is a natural overproduction of offspring by parents (commonly two parents produce many more offspring in their lifetime). 2. Genetic variability is found in all sexually reproducing organisms. 3. Genetic variability may be translated into physical variability (in morphology, physiology or behaviour). 4. The environment will normally favour one variant over another (we say the organism is better adapted). 5. The more favoured type will thrive while the less favoured type will be disadvantaged (we say the more favoured type has a selective advantage).

80

Natural Selection and Adaptive Change

6. By virtue of its more favourable circumstances the organism with the advantage will generally live longer (increased survivorship) and produce more offspring (a greater fecundity). The better adapted organism has increased fitness. 7. Genetic variability is heritable (hence parents can pass these favourable characteristics on to their offspring). 8. And, over time, the population will change as the better adapted forms make up a greater and greater percentage of the population. In the summation thus far, adaptive evolutionary change has resulted from the action of natural selection. This formulation though then leads to two further problems:  While all adaptive evolutionary change may result from natural selection; natural selection does not always result in evolutionary change.  If there is also a non-adaptive evolutionary change, such that two or more hereditary variants do not differ from one another in degree of adaptedness (i.e. fitness), then evolutionary change from one to the other cannot be a result of natural selection. In the later editions of Origin of Species Charles Darwin borrows Herbert Spencer’s expression ‘survival of the fittest’ to describe animals and plants that appeared to be best adapted to their environment. However, C. H. Waddington in 1960 suggested that the phrase ‘survival of the fittest’ might be tautological (repeating a meaning that has already been conveyed). And if we equate fitness with survivorship, then that is indeed true – we have ‘survival of the survivors’! However, in modern usage the term ‘fitness’ is a portmanteau word implying a general reproductive fitness including several components (fecundity, viability, survivorship, etc.) and, at best, we can call this a partial tautology.

Selection in Populations In an unchanging environment natural selection should theoretically result in a supremely adapted population of individuals. However, fluctuations in the surroundings will often produce a shift in the direction of selection while influx of new individuals might introduce new traits and therefore alter existing selection pressures. In such examples populations will be sorted according to new conditions and, after several generations, new population parameters can be described. Within a population natural selection can affect the frequency of a heritable trait in three or four different ways (Figure 4.2): 1. The type of selection that results in direct evolutionary change in a population over time is known as directional selection. This is most easily envisaged if one imagines selection acting on some continuous variable, like adult body size (which will probably be due to the additive effect of some gene loci). In directional selection one extreme phenotype is the fittest. 2. In the case of a normalising selection some intermediate size is optimal (‘selection for ordinariness’) – over time the extremes of the range will be reduced in frequency. Some authors use the term ‘stabilising selection’ for this mode.

Polymorphism

81

Figure 4.2 Diagrammatic representation of (a) stabilising, (b) directional and (c) disruptive

selection.

3. Here, the use of the term stabilising selection is reserved for describing selection that maintains the status quo, rather than resulting in elimination of the extremes. 4. The fourth mode is disruptive selection: two or more phenotypes are fitter than the intermediates between them, so the population becomes polymorphic. In the simplest case for body size, the largest and the smallest phenotypes are favoured.

Polymorphism Organisms vary genetically and phenotypically both within and between populations. They can exist in many forms; a feature we call polymorphism. The variation includes both quantitative (continuous) and discrete (discontinuous) characters; (incidentally, geneticists generally use the term character for a heritable feature, flower colour, etc., that varies between individuals in a population; each variant for that character, red or white flowers for example, is referred to as a trait). The investigation of polymorphism (presumed to have originated by disruptive selection) was regarded as particularly important in the development of the Synthetic Theory from the 1930s onwards. Before that time, laboratory geneticists had a picture of the nature of the genome as consisting of a series of mainly homozygous (identical) loci each with a pair of the ‘best’ alleles – the ‘wildtype’. A few loci, probably heterozygous, were occupied by ‘mutant’ alleles that were almost uniformly less fit

82

Natural Selection and Adaptive Change

than their respective wildtype (unless active directional selection was in play). These mutant alleles manifested themselves as the obvious abnormalities on which transmission genetics was principally founded. Notably this was the case with Drosophila, but nobody, at least in Western Europe or North America, had studied the nature of the Drosophila genome in wild populations. From the mid-1930s onwards, such studies were undertaken in California by Theodosius Dobzhansky (1900–1975), a Russian émigré whose most notable work, with a series of colleagues, was undertaken in the laboratory and the field on the species D. pseudoobscura. Dobzhansky discovered not only that the populations were polymorphic at many gene loci, but also that there was extensive chromosome polymorphism as well. It was a fortunate circumstance that several species of Drosophila had been used in early research, because in that genus it was subsequently discovered that in the salivary gland there were ‘giant chromosomes’ consisting of hundreds of copies of each single chromosome bound together and extended along their length. Each polytene or giant chromosome has a characteristic pattern of bands like a bar code, allowing each gene locus to be assigned a position relative to nearby stripes. Chromosome polymorphism is due to ‘mistakes’ at cell division, so that lengths of chromosome are inverted within the chromosome, or bits of chromatid separate in association with the ‘wrong’ partner chromatid. Dobzhansky found that these abnormal chromosomes were retained in wild populations not as abnormal freaks but at a high frequency, often in the heterozygous (or more correctly heterokaryotype) condition. When an abnormal chromosome is paired with a normal one, they go through all sorts of contortions in an’ attempt’ to pair corresponding loci. The resulting bizarre shapes can easily be seen in the giant chromosomes. Dobzhansky was also able to show from observations in the field and the laboratory that the differences between these polymorphic karyotypes were of adaptive significance. A karyotype is a diagram or photograph showing the organised sequence of chromosomes within a cell. One example here will suffice. In D. pseudoobscura there are several karyotypes of the third chromosome. Two are known as Standard (ST) and Chiricahua (CH). In the wild the relative frequencies of these two vary in systematic fashion throughout that period of the year when the flies are active. In the laboratory the two karyotypes were left to interbreed freely for several generations. It was found that at any given temperature the ratio between the two stabilised to a constant value, whatever the initial frequencies were. At 25 C these were about 70% ST and 30% CH; however, at 15 C they were equal at 50%. In the wild at first ST occurs at more than twice the rate of CH but declines from the beginning of June when the temperatures of high summer begin. Then ST increases rapidly, as better heat adapted, until the end of the season. The increase of CH from March to the beginning of June was suggested as being due to its advantage in sparser populations before the summer build-up in numbers. This alteration in frequency could be compared with selection on beak size and shape in the Darwin’s finch G. fortis described in Chapter 1. Both are examples of stabilising selection in the sense that we use the term: the maintenance of a range of frequencies,

Polymorphism

83

in the case of Drosophila a stable but seasonal polymorphism; in the case of Geospiza, the range of a continuous variable dictated by a less predictable external cause. During the same period in the middle of the twentieth century that Dobzhansky and his colleagues were investigating the hidden polymorphisms of Drosophila, E. B. Ford and a few associates and colleagues were studying visible polymorphisms as part of the research programme that Ford himself described as ‘Ecological Genetics’ (the title of his text on the subject in 1964). Ford’s definition of genetic polymorphism is as follows: Genetic polymorphism is the occurrence together in the same locality of two or more discontinuous forms of a species in such proportions that the rarest of them cannot be maintained merely by recurrent mutation.

Ford goes on to say ‘. . . Evidently it [the definition] excludes geographical races, as well as continuous variation controlled by polygenes, and falling within a curve of normal distribution, as with human height. It excludes also the suggestion of rare recessives, or heterozygous conditions eliminated by selection and maintained only by mutation pressure’. Differences between one generation and the next, as between spring and summer forms of some butterflies, are also excluded. It is obvious, a priori, that two kinds of genetic polymorphism could exist. If in a population some hereditary trait, in the simplest case due to a single gene, is replacing its ‘normal’ allele because of selection, then until the new allele has gone to fixation, both will be present in the population ‘in such proportions that the rarest of them cannot be maintained merely by recurrent mutation’. This is a transitional polymorphism due to directional selection. Stable polymorphism is exemplified by Dobzhansky’s work on Drosophila. Another highly variable polymorphic species is the common weed of wasteland, the dandelion, Taraxacum officinale complex. Dandelion is a globally widespread and exceptionally variable plant. There are more than 2000 subspecies of Taraxacum with 90% of them polyploid (chromosome number ranges from 16 to 48 with ploidy levels rising to hexaploid). Although exhibiting quite wide-ranging plasticity in body form, the plant is still considered a single species (though an aggregate group for practical purposes) and debate is still on-going as to whether the observed variation is due to phenotypic plasticity or genetic differentiation arising from multiple introductions across Europe and the Americas. It is apparently of recent speciation along with other apomictically (asexually) reproducing plants such as bramble (Rubus) and rose (Rosa). In T. officinale, virtually every floret produces a seed which is genetically identical to the mother so that mother–daughter lines form 'seed clones' (Richards, 1996) hence retaining the stable polymorphism. As mentioned in the previous chapter, several cases of polymorphism were investigated by the Oxford school of Ford and his associates (seeking adaptive explanations of visible polymorphism). A notable case was that of shell colour and banding pattern in the snail Cepaea nemoralis (and in its sister-species C. hortensis). For most of the twentieth century this has been the subject of investigation, and the investigation continues.

84

Natural Selection and Adaptive Change

Cepaea is an outcrossing hermaphrodite, so that after mating both partners produce eggs. The offspring can have shells of various background colours, inherited as an allelic series, and bracketed as brown (the dominant form), pink and yellow (with pale yellow as the recessive). This genetic locus is closely linked to a locus that determines whether bands occur on the shell. Without the recessive bands being present no banding can occur. Also, closely linked to these two loci is one that may or may not give rise to the presence of bands. Unlinked loci then determine the number of bands. These snails normally have a maximum of five, usually brownish-black bands that spiral down the shell from the apex to the lip of the aperture. A 5-banded specimen is recorded as 12345, if only band 3 is present (mid-banded), 00300, absence of bands 1 and 2 as 00345, and so on (totally unbanded = 0000). Spread bands refer to those that are wider than non-spread; 4 and 5 particularly can be so wide that they run into one another. Other loci control band colour, with rare variants, and body colour. Most populations are polymorphic for some of the possible variants. From the 1950s onward Cain and Sheppard and others studied Cepaea from a selection point of view. At first it was concluded that colour and banding pattern together acted as camouflage. Populations (relatively static and occupying a small area: